A window into the Early to mid-Cretaceous infrastructure of the

advertisement
J. metamorphic Geol., 2013, 31, 729–753
doi:10.1111/jmg.12042
A window into the Early to mid-Cretaceous infrastructure of the
Yukon-Tanana terrane recorded in multi-stage garnet of westcentral Yukon, Canada
R. D. STAPLES,1 H. D. GIBSON,1 R. G. BERMAN,2 J. J. RYAN3 AND M. COLPRON4
1
Department of Earth Sciences, Simon Fraser University, Burnaby, BC, V5A 1S6, Canada (rdstaple@sfu.ca)
2
Geological Survey of Canada, Ottawa, ON, K1A 0E9, Canada
3
Geological Survey of Canada, Vancouver, BC, V6B 5J3, Canada
4
Yukon Geological Survey, Whitehorse, YK, Y1A 2C6, Canada
ABSTRACT
Amphibolite facies metasedimentary schists within the Yukon-Tanana terrane in the northern Canadian Cordillera reveal a two-stage, polymetamorphic garnet growth history. In situ U-Th-Pb Sensitive
High Resolution Ion Microprobe dating of monazite provide timing constraints for the late stages of
garnet growth, deformation and subsequent decompression. Distinct textural and chemical growth
zoning domains, separated by a large chemical discontinuity, reveal two stages of garnet growth characterized in part by: (i) a syn-kinematic, inclusion-rich stage-1 garnet core; and (ii) an inclusion-poor,
stage-2 garnet rim that crystallized with syn- to post-kinematic staurolite and kyanite. Phase equilibria modelling of garnet molar and compositional isopleths suggest stage-1 garnet growth initiated at
~600 °C, 8 kbar along a clockwise P–T path. Growth of the compositionally distinct, grossular-rich,
pyrope-poor inner portion of the stage-2 overgrowth is interpreted to have initiated at higher pressure
and/or lower temperature than the stage-1 core along a separate P–T loop, culminating at peak P–T
conditions of ~650–680 °C and 9 kbar. Stage-2 metamorphism and the waning development of a
composite transposition foliation (ST) are dated at c. 118 Ma from monazite aligned parallel to ST,
and inclusions in syn- to post-ST staurolite and kyanite. Slightly younger ages (c. 112 Ma) are
obtained from Y-rich monazite that occurs within resorbed areas of both stage-1 and stage-2 garnet,
together with retrograde staurolite and plagioclase. The younger ages obtained from these texturally
and chemically distinct grains are interpreted, with the aid of phase equilibria calculations, to date
the growth of monazite from the breakdown of garnet during decompression at c. 112 Ma. Evidence
for continued near-isothermal decompression is provided by the presence of retrograde sillimanite,
and cordierite after staurolite, which indicates decompression below ~4–5 kbar prior to cooling below
~550 °C. As most other parts of the Yukon-Tanana terrane were exhumed to upper crustal levels in
the Early Jurassic, these data suggest this domain represents a tectonic window revealing a much
younger, high-grade tectono-metamorphic core (infrastructure) within the northern Cordilleran orogen. This window may be akin to extensional core complexes identified in east-central Alaska and in
the southeastern Canadian Cordillera.
Key words: in situ monazite geochronology; multi-stage garnet; P–T–t path; SHRIMP; YukonTanana terrane.
INTRODUCTION
The geodynamic evolution of orogenic belts is dictated to a large extent by the rheology of the lithosphere, which in turn is controlled, in part, through a
dynamic interplay between metamorphism and
deformation. Understanding of the evolution of orogenic belts is, therefore, significantly aided by establishing the P–T conditions and style of deformation
and metamorphism operating at various times and
locations throughout an orogen. However, the
diachronous and transitory nature of deformation
and metamorphism across an orogen prohibits the
© 2013 John Wiley & Sons Ltd
correlation of metamorphic and deformational events
solely on the basis of similarities in metamorphic
mineral assemblage and structural style (Williams,
1985). This issue is compounded in the cores of orogenic belts, which may experience multiple episodes
of deformation and metamorphism under nearly
identical conditions.
The advent of in situ U-Pb monazite geochronology allows monazite to be dated in its petrological
context with respect to metamorphic assemblages and
fabric elements, thereby providing a greater level of
confidence in establishing not only the age of metamorphism and the associated P–T conditions (Foster
729
730 R. D. STAPL ES ET AL .
et al., 2002, 2004; Gibson et al., 2004) but also the
timing of deformation fabrics (Williams & Jercinovic,
2002; Berman et al., 2005, in press). Additionally,
when coupled with analysis of garnet zoning patterns
and phase diagram modelling, in situ U-Pb monazite
dating can establish the timing of points along specific P–T paths (Berman et al., 2007a, 2010, in press;
Caddick et al., 2007; Gaidies et al., 2008). Applications of this method have also identified domains of
contrasting P–T–t histories in areas previously considered to be part of a single tectono-metamorphic
unit (Berman et al., 2007a; Horv
ath et al., 2010).
We apply similar techniques to amphibolite facies
rocks of the Yukon-Tanana terrane, which occupies
much of the metamorphic core of the orogen in the
northern Canadian Cordillera and easternmost
Alaska (Figs 1 & 2). Typical of the core zone of
many orogenic belts, the Yukon-Tanana terrane
rocks are poly-deformed and metamorphosed, with
metamorphic and deformational events recorded in
the Devono-Mississippian, Permo-Triassic, Jurassic
and Cretaceous (see Berman et al., 2007a for a
review). Further complicating matters, mapping and
P–T-t–D data (Berman et al., 2007a) reveal that the
Arctic Ocean
style, pattern and grade of metamorphism and deformation was nearly identical in each of the Permo-Triassic, Jurassic and Cretaceous tectono-metamorphic
events.
Difficulty in identifying distinct tectono-metamorphic domains within the Yukon-Tanana terrane, and
deciphering discrete deformational and metamorphic
events within an individual domain, are overcome
herein by a detailed analysis of textural relationships
and monazite and garnet chemistry. The timing of
deformation and a segment of an individual P–T
path are determined by in situ Sensitive High Resolution Ion Microprobe (SHRIMP) monazite geochronology, with monazite growth texturally and
chemically linked to deformation fabrics as well as
metamorphic porphyroblasts and their modelled P–T
stability conditions. These data, together with the
data of Berman et al. (2007a), elucidate a high-pressure Cretaceous tectono-metamorphic domain (Australia Mountain domain) that is distinct from an
adjacent domain affected by Permo-Triassic and
Early Jurassic events, and exhumed to upper crustal
levels in the Early Jurassic. In this study, we better
constrain the timing, nature and regional extent of
Cretaceous metamorphism and decompression in
west-central Yukon, and elucidate the structural-thermal architecture of the northern Cordilleran orogen
in the Cretaceous.
GEOLOGICAL BACKGROUND
YT
AK
YT
NWT
NAm
Ti
nt
in
a
lim
it
eastern
Nab
of
D
pNAm
en
Fig. 2
al W
i
Fa
ul
t
t
ul
Fa
WR
n
illera
Cord
NAm
D
Northern Cordilleran orogen
N
YT
QN
NAb
WR
CA
YT
Wh
Pacific
Ocean
0
200
YT
deformation
ST
KS
AX
SM
BC
CC
SM
km
Figure 1. Simplified terrane map of the northern Canadian
Cordillera showing location of Fig. 2, and the study area,
within the northern portion of Yukon-Tanana terrane
(modified from Colpron et al., 2006). D, Dawson; Wh,
Whitehorse. Lithotectonic terrane abbreviations: AX,
Alexander; CA, Cassiar; CC, Cache Creek; KS, Kluane Schist;
NAb, North American basinal strata; NAm, North American
platformal strata; pNAm, parautochthonous North American
continental margin; SM, Slide Mountain; ST, Stikinia; YT,
Yukon-Tanana; W, Windy; WR, Wrangellia.
The Yukon-Tanana terrane, together with Stikinia,
Quesnellia and the oceanic Slide Mountain terrane,
represent late Palaeozoic to Mesozoic arc and backarc systems that lay outboard of the previously
thinned, western ancestral North American margin
(Colpron et al., 2007). The timing and nature of
accretion of the above terranes have been well documented in both the northern and the southern Canadian Cordillera (e.g. Monger et al., 1982; Nelson
et al., 2006; Colpron et al., 2007; Beranek & Mortensen, 2011). In the Mesozoic, the orogen grew as a
consequence of protracted compression and crustal
thickening during arc-continent collisions (e.g., Berman et al., 2007a; Gibson et al., 2008) as the North
American craton moved westward, converging with
its offshore subduction zone (Monger & Price, 2002).
In the southern Canadian Cordillera, a significant
body of geochronological data from the metamorphic
core of the orogen has revealed that despite similarities in regional metamorphic grade and deformation
patterns, both were strongly diachronous, younging
systematically with increasing structural depth (Parrish, 1995). Rocks presently in the upper exposed
structural levels were buried, heated and exhumed in
the Jurassic (Murphy et al., 1995; Colpron et al.,
1996; Gibson et al., 2005), while structurally deeper
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 731
b
TERTIARY AND YOUNGER
Volcanics
LATE CRETACEOUS
Carmacks Group volcanics
b
LATE EARLY CRETACEOUS
a
Indian River formation
c
EARLY JURASSIC
60°N
Granitoid intrusions
120°W
139°W
PERMIAN
Sulphur Creek orthogneiss
e
Klondike schist
d
MISSISSIPPIAN TO PERMIAN
Ti
nt
in
a
Ultramafic
fa
ul
t
49°N
Ultramafic
Australia
Mountain
domain
63°40'N
DEVONIAN TO MISSISSIPPIAN
Granitoid
Selwyn Basin
Reid Lakes volcanoplutonic complex
Nasina assemblage
B
2&3
ult
stra
r fa
ive
ek
fau
rt R
Cre
Australia
Mountain
Amphibolite
LATE DEVONIAN AND OLDER
Marble
Snowcap assemblage
Ste
lt
wa
lia
63°40'N
Au
1&4
Marble
Undefined fault
Thrust fault
Normal fault
Dextral strike-slip fault
Geological contacts
C
W
illo
w
139°W
1
Sample location
La
ke
fa
ul
t
63°20'N
63°20'N
A
138°W
0
10
20 km
Figure 2. (a) Simplified geology for portions of Stewart River and McQuesten map areas showing the distribution of sample
locations and the approximate boundary of the Early to mid-Cretaceous Australia Mountain metamorphic domain bounded by the
Australia Creek, Stewart River and Tintina faults. Modified from Gordey & Ryan, 2005; Ryan et al., 2010. (b) Map showing the
distribution of rocks at mid- to lower crustal levels in the Early to mid-Cretaceous in the BC, Yukon and Alaskan Cordillera.
Location abbreviations and references: a, this study and Berman et al. (2007a); b, east-central Alaska (Pavlis et al., 1993; DuselBacon et al., 2002); c, Finlayson region (Murphy, 2004); d, Shushwap complex (Parrish, 1995; Crowley et al., 2000; Gibson et al.,
2008); e, Prince Rupert area (Wolf et al., 2010).
rocks were progressively buried and heated from at
least Cretaceous to earliest Eocene (Carr, 1991; Parrish, 1995; Crowley & Parrish, 1999; Gibson et al.,
1999, 2005; Crowley et al., 2000). The deepest structural levels within the core of the orogen, which
includes autochthonous and parautochthonous North
American crust, were largely exhumed by extensional
shear zones in the early Eocene (Parrish et al., 1988;
Brown et al., 2012), marking a shift to a transtensional tectonic setting.
The deformational and metamorphic history of the
Yukon-Tanana terrane within the northern Cordillera
differs from the southeastern Canadian Cordillera in
© 2013 John Wiley & Sons Ltd
that the main phase of metamorphism (amphibolite
facies) and deformation appears to have occurred
earlier, in the Late Permian to Early Triassic (Berman et al., 2007a; Beranek & Mortensen, 2011).
There was an ~8 kbar metamorphic overprint in the
Early Jurassic before much of the metamorphic core
of the northern Cordilleran orogen was rapidly
exhumed in the Early to Middle Jurassic (Hansen
et al., 1991; Stevens et al., 1993; Johnston et al.,
1996; Berman et al., 2007a).
In east-central Alaska, the allochthonous YukonTanana terrane is structurally underlain by a large area
of metamorphic rocks interpreted by Dusel-Bacon
732 R. D. STAPL ES ET AL .
et al. (1995, 2002) to be part of the parautochthonous western continental margin of ancestral North
America. In the pre-latest Triassic (>212 Ma), the
uppermost structural levels of the allochthonous
Yukon-Tanana terrane were affected by northeastdirected shear at 8–12 kbar (Dusel-Bacon et al.,
1995). By Early Jurassic (>188 Ma), both the lower
structural levels of the allochthonous Yukon-Tanana
terrane and the underlying parautochthonous rocks
were deformed and metamorphosed at amphibolite facies conditions (7–12 kbar) during northwestdirected contraction and imbrication (Dusel-Bacon
et al., 1995). Early Jurassic 40Ar/39Ar cooling ages
are interpreted by Hansen & Dusel-Bacon (1998),
and Dusel-Bacon et al. (2002) to date the time of
cooling of the upper plate following the aforementioned northwest-directed contraction that emplaced
the allochthonous rocks of Yukon-Tanana terrane
onto the parautochthonous continental margin. The
lower plate was subsequently exhumed by southeastdirected crustal extension as recorded in c. 135–
110 Ma 40Ar/39Ar cooling ages in parautochtonous
continental margin rocks (Pavlis et al., 1993; Hansen
& Dusel-Bacon, 1998; Dusel-Bacon et al., 2002).
Exhumation of ductily deformed amphibolite facies
rocks in the Early to mid-Cretaceous is also recorded
<200 km to the east of the Cretaceous domain of
east-central Alaska, in the Australia Mountain area
of west-central Yukon. Although the timing of deformation and metamorphism appears to be older in the
northern Cordillera, this diachroneity of tectonism,
which may be a function of structural level and position within the orogenic wedge, suggests similarity
with what is recorded in the southern Canadian
Cordillera.
Australia mountain area
The Yukon-Tanana terrane has been dissected and
offset ~430 km by the Tintina fault, an Eocene dextral strike-slip fault (Gabrielse et al., 2006). The study
area is located at Australia Mountain (Figs 1 & 2),
immediately southwest of the Tintina fault, and is
underlain by highly deformed psammite, semi-pelite
and quartzite with lesser amounts of pelite, mafic volcanic and ultramafic rocks that have been metamorphosed to amphibolite facies. These rocks are widely
accepted to be part of Yukon-Tanana terrane (eg.
Mortensen, 1990; Gordey & Ryan, 2005; Ruks et al.,
2006; Berman et al., 2007a), and we treat them as
such here. However, as is elaborated in the discussion
section, our data allow that the rocks could equally
represent parautochthonous North American continental margin rocks that are structurally juxtaposed
with Yukon-Tanana terrane similar to what is
described by Dusel-Bacon et al. (2002) and DuselBacon et al. (2006) in east-central Alaska. West and
south of the Australia Mountain area are polydeformed and transposed amphibolite facies Palaeozoic
rocks of the Yukon-Tanana terrane, which are
intruded by weakly to undeformed plutons of Triassic, Jurassic, Cretaceous and Eocene age (Gordey &
Ryan, 2005). These are overlain by Cretaceous and
Paleocene volcanic and sedimentary rocks, and rare
Quaternary basalt (Gordey & Ryan, 2005).
The main episode of deformation to the west and
south of Australia Mountain, and generally considered
representative for the Yukon-Tanana terrane in general, produced a regional transposition foliation that is
interpreted by Berman et al. (2007a) and Beranek &
Mortensen (2011) to be bracketed between c. 260 and
253 Ma. This constraint is based upon the ages
obtained from the strongly foliated Sulphur Creek orthogneiss (c. 260 Ma), as well as the undeformed Jim
Creek pluton (253 Ma), which cuts across the transposition foliation in the Devono-Mississippian Nasina
assemblage (Beranek & Mortensen, 2011). A c.
239 Ma monazite included within garnet was interpreted by Berman et al. (2007a) to constrain the latter
stages of a Late Permian to Middle Triassic tectonometamorphic event that was initiated at low pressure
and culminated with the growth of garnet at ~9 kbar
and 600 °C. A subsequent episode of metamorphism is
recorded in c. 195–187 Ma monazite inclusions within
garnet and kyanite porphyroblasts, which are interpreted by Berman et al. (2007a) to date an Early
Jurassic metamorphic event with peak conditions of
~8 kbar and 600 °C. Unlike the Permo-Triassic event,
this Early Jurassic metamorphism was not accompanied by significant fabric development, rather Berman
et al. (2007a) suggested the strain was partitioned heterogeneously into regional high-strain zones. West of
Australia Mountain, amphibolite facies rocks of the
Snowcap assemblage and structurally lower level
Permian rocks at greenschist facies were interpreted by
Mackenzie & Craw (2012) to be thrust imbricated
along localized ductile and brittle Jurassic shear zones
that were active at greenschist facies conditions.
The amphibolite facies rocks west and south of the
Australia Mountain area yield Early to Middle Jurassic K-Ar cooling ages (Hunt & Roddick, 1992), interpreted to record widespread exhumation and cooling
below ~300 °C. In contrast, strongly deformed,
amphibolite facies rocks at Australia Mountain are
juxtaposed to the southeast against essentially undeformed and unmetamorphosed rocks of the Mississippian Reid Lakes volcano-plutonic complex
(Colpron & Ryan, 2010). The preservation of Mississippian 40Ar/39Ar cooling ages throughout the Reid
Lakes complex (Knight et al., 2013) indicates it was
not significantly buried or heated during the Permian
and Jurassic tectonometamorphic events. This is also
consistent with the lack of observed deformation and
metamorphism within the complex. The boundary
between the Australia Mountain domain and these
adjacent distinct domains are inferred as normal
faults, herein named the Australia Creek and Stewart
River faults (Fig. 2).
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 733
At Australia Mountain, the rocks are characterized by a strong, penetrative foliation that is parallel
to the axial planes of tight to isoclinal folds. Primary compositional layering in metasedimentary
rocks and a pre-existing foliation can be traced
around the closures of these tight to isoclinal folds
indicating that the folds are at least F2 structures.
The presence of rootless intrafolial isoclinal folds
suggests this foliation is likely a composite transposition foliation in the sense of Williams (1983) and
Tobisch & Paterson (1988). Following Williams &
Compagnoni (1983), we have notated this composite
transposition foliation as ST. ST has in turn been
deformed by at least two later episodes of folding,
herein denoted FT + 1 and FT + 2. The FT + 1 folds
are subhorizontal to gently plunging, steeply to
moderately inclined, tight asymmetric chevron folds
with more open parasitic crenulations. The FT + 1
folds lack an axial planar foliation, and show no
signs of development of a crenulation cleavage. The
latest phase of folding, FT + 2, consists of subhorizontal, upright gentle to open folds. The metamorphic grade, style, orientation and pattern of
successive phases of deformation within this Cretaceous domain are strikingly similar to those found
in the surrounding Permo-Triassic and Jurassic
metamorphic domain (Ryan et al., 2003). Through
the application of high resolution in situ U-Pb monazite dating, linked to the P–T history and deformation fabrics, we herein shed light on the unique
Early to mid-Cretaceous metamorphic, deformational and exhumation history for this portion of
the Yukon-Tanana terrane.
PETROLOGICAL METHODS
Thermobarometry
Mineral compositions were quantitatively analyzed
using a fully automated CAMECA SX-50 instrument,
operating in wavelength-dispersion mode with the following operating conditions: excitation voltage,
15 kV; beam current, 20 nA (10 nA for mica); peak
count time, 20 s (40 s for F, Cl); background counttime, 10 s (20 s for F, Cl); spot diameter, 5 lm. Quantitative data were obtained for garnet, biotite, plagioclase and muscovite (Table 1). Temperatures and
pressures were calculated using the winTWQ program
(version 2.32; Berman, 2007), which uses internally
consistent thermodynamic data for end-members and
mixing properties to calculate the average P–T values
from the intersections among the following independent set of equilibria (mineral abbreviations after
Kretz, 1983):
Alm þ Phl ¼ Pyp þ Ann
(1)
Grs þ 2Ky þ Qtz ¼ 3An
(2)
3East þ 2Grs þ Pyp þ 6Qtz ¼ 3Phl þ 6An
(3)
© 2013 John Wiley & Sons Ltd
Alm þ Grs þ Ms ¼ 3An þ Ann
(4)
The winTWQ software incorporates solution solid
solution models for garnet and biotite (Berman,
2007), as well as for plagioclase (Fuhrman & Lindsley, 1988) and muscovite (Chatterjee & Froese, 1975).
For samples 1 and 2 with sodic-rich plagioclase, the
aluminium-avoidance plagioclase model of Aranovich
(1991) was used. This model yields lower pressures
that are generally more consistent with independent
estimates (Aranovich, 1991). Absolute errors of thermobarometric data are considered to be ~ 50 °C
and 1 kbar (Berman, 1991), with appreciably smaller
errors associated with relative differences between
samples.
Phase diagram calculations
Isochemical phase diagram sections were calculated
in the system MnO-Na2O–CaO–K2O–FeO–MgO–
Al2O3–SiO2–H2O–TiO2-Fe2O3 with the program
Domino (De Capitani & Brown, 1987; De Capitani
& Petrakakis, 2010; http://titan.minpet.unibas.ch/
minpet/theriak/theruser.html, version 01.08.09), and
the internally consistent thermodynamic data set of
Holland & Powell (1998); data set tcds55, created on
22 November 2003. Activity-composition models
were used for the following phases: muscovite,
excluding the margarite component (Coggon & Holland, 2002), feldspar (Holland & Powell, 2003), garnet and biotite (White et al., 2007), ilmenite-hematite
and magnetite-ulvospinel (White et al., 2000), chlorite, cordierite, chloritoid, staurolite and epidote
(Holland & Powell, 1998). All other phases were treated as pure. Garnet, biotite, staurolite, chloritoid,
chlorite and cordierite were extended to the Mnbearing system as outlined in Tinkham et al. (2001).
A pure H2O fluid was considered in excess in all
calculations. Additionally, excess SiO2 was added for
sample 1 calculations. Silicate melt was not included
in the calculations. Bulk rock compositions (Table 2a)
were determined by whole-rock XRF analysis from
thin section offcuts. A nominal amount of Fe2O3
(0.01 mol.% Fe2O3) was added to the system to be
consistent with the garnet and biotite solution models
of White et al. (2007), which incorporate Fe2O3. A
small amount of Fe2O3 is consistent with ilmenite as
the only Fe-bearing oxide, suggesting these rocks are
fairly reduced.
The effective bulk composition of a rock changes
with P–T as material is progressively fractionated
within the cores of mineral grains with slow intracrystalline diffusion rates (St€
uwe, 1997). Evidence for
such slow diffusivity, and a changing effective bulk
composition, is seen in the preservation of zoned
minerals, in particular garnet. Phase diagrams calculated from the bulk whole-rock analysis are therefore
only strictly correct at the P–T conditions during the
734 R. D. STAPL ES ET AL .
Table 1. Microprobe analyses used in P–T calculations.
Sample 1 (09RS190A1)
Sample 2 (09RS172B1)
Sample 3 (09RS171A1)
Mineral
Position
Grt-2
Rim
Pl
Bt
Grt-1
Core
Grt-2
Rim
Pl
Bt
Grt-1
Core
Grt-2
Rim
Pl
Bt
Ms
wt% oxides
SiO2
TiO2
Al2O3
FeO*
MgO
MnO
CaO
Na2O
K2 O
Total
37.13
0.00
21.94
32.79
4.93
0.55
2.18
0.00
n.d.
99.53
63.13
n.d
22.59
0.35
0.00
0.02
3.68
9.21
0.11
99.09
37.32
1.56
17.94
17.40
12.50
0.05
0.00
0.45
8.61
96.32
36.74
0.06
21.20
35.72
2.53
1.02
2.28
0.02
n.d.
99.58
36.52
0.03
21.46
34.46
3.20
0.50
3.59
0.01
n.d.
99.81
61.61
n.d.
24.08
0.10
0.00
0.02
5.36
8.42
0.09
99.67
35.50
1.84
18.61
20.53
9.35
0.03
0.00
0.27
8.99
95.70
37.03
0.07
21.43
33.31
1.23
2.08
5.53
0.00
n.d.
100.72
36.90
0.00
21.68
30.75
2.63
4.35
3.93
0.03
n.d.
100.28
61.22
n.d.
24.90
0.18
0.00
0.00
6.09
7.84
0.19
100.42
35.30
2.26
20.14
21.21
7.72
0.25
0.00
0.30
9.02
96.57
44.76
0.37
35.82
1.17
0.48
0.00
0.04
0.87
9.83
93.69
Cations
Oxygen
Si
Ti
Al
Fea
Mn
Mg
Ca
Na
K
12
2.96
0.00
2.06
2.18
0.04
0.59
0.19
0.00
n.d.
8
2.82
n.d.
1.19
0.01
0.00
0.00
0.18
0.80
0.01
11
2.78
0.09
1.57
1.08
0.00
1.39
0.00
0.06
0.82
12
2.98
0.00
2.03
2.42
0.07
0.31
0.20
0.00
n.d.
12
2.94
0.00
2.04
2.32
0.03
0.38
0.31
0.00
n.d.
8
2.74
n.d.
1.26
0.00
0.00
0.00
0.26
0.73
0.01
11
2.72
0.11
1.68
1.31
0.00
1.07
0.00
0.04
0.88
12
2.97
0.00
2.03
2.24
0.14
0.15
0.48
0
n.d.
12
2.96
0.00
2.05
2.06
0.30
0.31
0.34
0.00
n.d.
8
2.72
n.d.
1.30
0.01
0.00
0.00
0.29
0.67
0.01
11
2.68
0.13
1.80
1.34
0.02
0.87
0.00
0.04
0.87
11
3.03
0.02
2.86
0.07
0.00
0.05
0.00
0.11
0.85
0.44
0.89
0.81
0.10
0.07
0.02
0.86
0.76
0.13
0.10
0.01
0.55
0.94
0.75
0.05
0.16
0.05
0.87
0.69
0.10
0.11
0.10
Fe/(Fe + Mg)
XAlm
XPyp
XGrs
XSps
XAn
a
0.79
0.73
0.20
0.06
0.01
0.26
0.26
.
Table 2. (a) Representative bulk-rock geochemical analyses (wt
%). (b) Normalized effective bulk composition for the growth
of stage-2 garnet (wt%).
(a)
SiO2
TiO2
Al2O3
FeOa
MgO
MnO
CaO
Na2O
K2O
LOI
Total
(b)
SiO2
TiO2
Al2O3
FeOa
MgO
MnO
CaO
Na2O
K2O
LOI
Total
Grt-1 modea
0.30
2+
All Fe assumed to be Fe
Sample
0.61
1
2
3
51.20
0.90
21.40
10.15
3.01
0.12
2.37
5.72
1.54
0.40
96.81
43.40
1.30
30.50
11.25
3.23
0.10
1.37
2.44
1.92
1.97
97.48
42.80
1.06
23.00
16.90
2.38
0.92
4.00
2.20
2.89
0.40
96.55
54.43
1.00
22.34
8.37
2.83
0.03
2.48
6.36
1.71
0.44
100.00
7
44.97
1.38
31.72
10.50
3.29
0.07
1.35
2.59
2.04
2.09
100.00
3.4
45.53
1.21
24.31
15.40
2.51
0.89
3.90
2.51
3.29
0.46
100.00
9
Compositions in Table 2b represent the effective bulk composition during growth of
stage-2 garnet overgrowths (Grt-2) following the fractionation of components into the
core of stage-1 garnet (Grt-1). The effective bulk compositions were calculated by
subtracting the average composition of stage-1 garnet (Grt-1) from the whole-rock
XRF analysis, with the modal amount of Grt-1 estimated from image analysis of thin
sections.
a
Modal amount of Grt-1 subracted in the calculation of the effective bulk composition.
initial stage of garnet growth. Phase diagrams and
compositional isopleths calculated from the wholerock analysis were therefore used to estimate the prograde P–T path during the early stages of garnet
growth. Isochemical phase diagrams, which are more
suitable for modelling peak and retrograde metamorphic conditions, were constructed from an effective
bulk composition calculated by subtracting the average composition of stage-1 garnet from the whole-rock
XRF analysis. Whole-rock analyses and calculated
effective bulk compositions used in the construction of
phase diagrams are listed in Table 2a,b. There is no
evidence to suggest that the bulk composition was
further modified by open system behaviour. For
instance, the absence of evidence for partial melt
suggests the bulk composition was not modified by
melt loss. Furthermore, the mineralogical, chemical
and textural features suggest the rock was not
affected by fluid alteration following regional metamorphism.
Some discrepancy in P–T estimates can be
expected to arise from the different thermodynamic
data utilized in the phase equilibria modelling and
thermobarometric calculations. Nonetheless, the winTWQ program was utilized because its calibration of
equilibrium (3), a useful barometer for the aluminosilicate- and muscovite-absent rocks, is based on
experimental data directly constraining eastonite
properties (Berman et al., 2007b).
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 735
PETROGRAPHY AND MINERAL CHEMISTRY
Four samples from Australia Mountain were selected
for thermobarometric and geochronological analysis.
Metamorphic mineral assemblages and coordinates
of the sample localities are provided in Table 3. The
investigated samples are garnet-bearing psammitic
and pelitic schists characterized by garnet porphyroblasts that typically show two distinct domains: a resorbed, inclusion-rich darker coloured core (Grt-1)
and a lighter, inclusion-poor, euhedral overgrowth
(Grt-2) (Fig. 3a,c–e). A penetrative foliation (ST) is
generally defined within psammitic schists by quartz
ribbons, elongated plagioclase and micaceous-rich
layers. In some pelitic schists, staurolite and kyanite
porphyroblasts are consistently aligned parallel to ST,
and exhibit kinked, bent and sweeping extinction
indicative of syn-kinematic growth with respect to ST
(Fig. 3b,f). However, in other semipelitic schists staurolite and kyanite have a weak preferred orientation
and may, or may not, contain intracrystalline evidence for deformation. Many other staurolite and
kyanite grains in this same rock are randomly oriented with homogenous extinction, showing no evidence of strain. Staurolite and kyanite growth in the
Australia Mountain area is therefore interpreted as
syn- to post-kinematic with respect to ST.
Grt-1 occasionally contains a sigmoidal inclusiontrail (Si) that in rare samples is continuous with the
external transposition foliation (ST) (Fig. 3a), suggestive of early growth under conditions of non-coaxial
shearing. By contrast, Grt-2 has fewer primary inclusions, is typically subhedral to euhedral, and has
well-developed crystal faces modified by late-stage
resorption (Fig. 3c,d). Partial breakdown of Grt-2
rims to randomly oriented mats of biotite masks the
relationship of the foliation to Grt-2. Where Grt-2
has not been resorbed, it is characterized by straight,
euhedral crystal faces that appear to truncate ST
(Fig. 3c,d), indicative of post-kinematic growth.
The boundary between Grt-1 and -2 is consistently
marked by a large chemical discontinuity in all samples (Fig. 4). Grt-1 generally has low Fe/(Fe + Mg)
and grossular contents that both decrease rimward in
some samples. The inner portion of Grt-2 generally
has markedly higher grossular, and lower almandine
and pyrope than Grt-1. Fe/(Fe + Mg) and spessartine
decrease towards the outer rim, characteristic of prograde growth zoning (Tracy et al., 1976; Spear et al.,
1990; Florence & Spear, 1993). A slight kick-up in
spessartine and Fe/(Fe + Mg) at the outer 100 lm of
Grt-2 adjacent to both plagioclase and biotite is
attributed to retrograde garnet resorption (Kohn &
Spear, 2000). Biotite grains immediately adjacent to
garnet have the highest Fe/(Fe + Mg) values. Otherwise, biotite grains >300 lm from garnet are 0.05 to
0.07 lower in Fe/(Fe + Mg) and have very little compositional variation (<0.02). Matrix plagioclase has a
variable composition, with the lowest anorthite contents (XAn = 0.17–0.20) consistently within the cores
of matrix grains of all samples, and the highest anorthite content (XAn = 0.25–0.34) within the rims of
matrix grains immediately adjacent to strongly resorbed portions of garnet (XAn = 0.30).
Sample 1
Sample 1 is an Ilm-St-Grt-Bt-Pl-Qtz-bearing psammitic schist. It shares the textural and chemical features
described above, but shows additional complexities
and a unique textural relationship with monazite that
warrant more detailed description. Portions of Grt-1,
the outer portion of Grt-2, as well as the interface
between them have been partially replaced by plagioclase, fine-grained euhedral staurolite and Y-rich monazite (Fig. 3c,d), forming an island (Grt-1) and atoll
(Grt-2) structure. Staurolite and monazite do not
occur as inclusions in either Grt-1 or Grt-2. In addition
to the garnet chemical zoning described above, it is
noteworthy that garnet has the same composition at its
interface with the replacement products, St-Pl-Mnz, at
the Grt-1 rim, Grt-2 inner rim, and Grt-2 outer rim
(Fig. 4). Skeletal staurolite also occurs within the
matrix of sample 1 where it is aligned parallel to the
relict foliation and is deflected around garnet (Fig. 3d).
The pre- to syn-kinematic nature of this matrix staurolite, together with its resorbed/skeletal texture, suggests
it is an earlier generation than the euhedral, retrograde
Table 3. Summary of thermobarometric and geochronological data.
UTM (zone 7
NAD 83)
Stage-2 Grt rim peak
thermobarometry
Field no.
Easting
Northing
Main assemblage
1
09RS190A1
642809
7056955
St-Grt-Bt-Pl-Qtzd
2
3
4
09RS172B1
09RS171A1
09RS188A2
641621
641101
642122
7063232
7063420
7057171
Ky-St-Crd-Grt-Bt Qtz-Pl
Ms-Bt-Grt-Pl-Qtz
Sil-Grt-Crd-St-Ky- Bt-Pl-Qtz
Sample no.
Estimated uncertainties are 1 kbar and 50 °C (Berman, 1991).
Equilibria used to derive listed P–T values (pd = phase diagram constraints.
All ages determined by in situ SHRIMP analysis of monazite. All ages represent weighted mean
cussed in text.
d
Minerals listed from least to most abundant.
Pa (kbar)
T (°C)
8.8
8
8.9
9–10
–
650
650
650
680
–
Eq.b
1, 3
pd
1, 2, 3
1, 3, 4
Agec (Ma)
120
112
118
117
3
6
2
1
Interpretation
Prograde M2, synM2 decompression
Prograde M2, synPrograde M2, synPrograde M2, syn-
to post-ST
to post-ST
to post-ST
to post-ST
a
b
c
© 2013 John Wiley & Sons Ltd
206
Pb/238U ages (2 sigma errors) of texturally and chemically similar monazite grains dis-
736 R. D. STAPL ES ET AL .
staurolite that occurs within resorbed portions of
garnet.
Sample 2
Sample 2 is a weakly foliated Ilm-Ky-St-Crd-Grt-BtQtz-Pl-bearing semipelitic schist. Staurolite and kyanite have both a weak preferred orientation with
undulatory extinction and are also randomly oriented
with homogenous extinction. Staurolite, and less
commonly kyanite, is rimmed by cordierite that contains numerous inclusions of fine-grained biotite.
Garnet porphyroblasts are <8.0 mm in diameter and
commonly have an atoll texture characterized by a
subhedral garnet atoll (Grt-2) surrounding an interior
of quartz, biotite and plagioclase. Other garnet porphyroblasts are present which show no chemical indication of a multi-stage growth history. These
porphyroblasts are interpreted as stage-2 garnet from
their similarity in texture, composition and zoning
pattern to stage-2 overgrowths.
Sample 3
Sample 3 is a garnet-rich Ilm-Ms-Bt-Grt-Pl-Qtz-bearing psammitic schist. A penetrative schistose foliation
is defined at outcrop and hand sample scale by elongate feldspar grains, and aligned biotite and muscovite
that are deflected around garnet porphyroblasts. At
thin-section scale, mica is randomly oriented, possibly
due to a late episode of static recrystallization that
masks the foliation and its timing of development relative to garnet growth. However, the presence of quartz
and plagioclase pressure shadows aligned parallel to
the foliation on opposite sides of euhedral garnet porphyroblasts indicates that the latter stages of garnet
growth were synchronous with deformation. Muscovite,
which is present only in sample 3, has a slight phengitic
content (Si pfu = 3.1–3.2), generally non-detectable Ca,
and Na/(Na + K + Ca) values of 0.08–0.10.
Sample 4
Sample 4 is a Rt-Sil-Grt-Crd-St-Ky-Bt-Pl-Qtz-bearing
metapelite. Garnet occurs as rare, small (<1 mm),
heavily resorbed grains both in the matrix and as
inclusions within kyanite. Garnet that occurs both in
the matrix and included in kyanite is partly replaced
by biotite and plagioclase. Staurolite and kyanite
contain kinked and sweeping extinction and are consistently aligned parallel to ST, indicative of syn-kinematic growth with respect to ST (Fig. 3b,f). Trains of
elongated rutile grains are aligned parallel to, and
help define, the ST foliation. Cordierite appears texturally late, forming replacement rims around staurolite, and to a lesser extent around kyanite, and occurs
throughout the matrix with numerous inclusions of
biotite. A small amount of fibrolitic sillimanite is
present in small patches of radiating fibres (Fig. 3f).
PARAGENETIC INTERPRETATION OF TEXTURAL
AND CHEMICAL RELATIONS IN GARNET
Sample 3 is the only sample which records pronounced chemical zoning of the inclusion-rich garnet
core, with rimward decreasing XSps, XGrs and Fe/
(Fe + Mg), and increasing XAlm and XPyp characteristic of prograde growth zoning produced along a
clockwise P–T path (Tracy et al., 1976; Spear et al.,
1990; Florence & Spear, 1993). However, the sharp
compositional gradient/discontinuity between Grt-1
and Grt-2 (Fig. 4), with the latter characterized by
an increase in Fe/(Fe + Mg) and a decrease in XPyp,
does not conform to a P–T path of increasing temperature required for garnet growth (Figs 5 & 6).
Rather, the sharp compositional gradient between
Grt-1 and -2 is interpreted as a zone of incomplete
diffusional re-equilibration that developed following
growth of the chemically distinct garnet overgrowth
(Grt-2). The arrangement of XGrs isopleths for sample 2 (Fig. 5) suggests growth of the grossular-rich
Grt-2 initiated at significantly higher pressure and/or
lower temperature than Grt-1.
Within the core of Grt-2, the decreasing Fe/
(Fe + Mg), XSps and XGrs values away from the zone
of diffusional re-equilibration towards the rim is
interpreted as a growth zoning pattern similar to
Grt-1, but one that developed during a subsequent
metamorphic event. Enrichment of spessartine and
Fe/(Fe + Mg) at the extreme Grt-2 rim, and Feenriched biotite immediately adjacent to resorbed
portions of garnet, are features characteristic of
Figure 3. Photomicrographs of samples 1, 2 and 4. (a) Garnet porphyroblast of sample 1 consisting of: (i) a syn-kinematic garnet
core (Grt-1), which hosts a sigmoidal inclusion trail (Si) that is continuous with the external transposition foliation (ST), and (ii) an
inclusion-poor overgrowth (Grt-2) on the top and bottom of the porphyroblast. (b) Sample 4, hand sample with elongate synkinematic staurolite (St) and kyanite (Ky) porphyroblasts consistently aligned parallel to the transposition foliation (ST). (c) Garnet
porphyroblast from sample 1 consisting of an inclusion-rich core (Grt-1) surrounded by an annulus of euhedral staurolite and
plagioclase, and an inclusion-poor garnet atoll (Grt-2). Both the garnet core and atoll from sample 1 are resorbed and replaced by
staurolite and plagioclase. The microprobe traverse displayed in Fig. 4a is shown in blue. (d) Garnet core and atoll of sample 1
replaced by euhedral staurolite and plagioclase. Skeletal matrix staurolite aligned parallel to the transposition foliation (ST) is
deflected around the garnet porphyroblasts, and is interpreted as an earlier generation than that which replaced garnet. (e) Garnet
porphyroblast from sample 2, which consists of an inclusion-rich garnet core (Grt-1), and, with the exception of ilmenite (black
inclusion), a relatively inclusion-poor overgrowth (Grt-2). The blue line across garnet shows the location of the microprobe
traverse displayed in Fig. 4b. (f) Syn-kinematic porphyroblasts of staurolite and kyanite aligned parallel to ST in sample 4.
Fibrolitic sillimanite (Sil) occurs as small radiating fibres. Note: All mineral abbreviations used in figures are after Kretz (1983).
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 737
ST+1
(a)
(b)
Grt-2
ST
Si
ST
St
Grt-1
Ky
1 mm
1 mm
Grt-2
(c)
(d)
ST
St
St
Grt-1
Grt-2
St
Grt-1
St
St
Grt-2
1 mm
(e)
1 mm
(f)
Ky
St
Grt-1
Grt-2
ST
Ky
St
St
1 mm
Sil
© 2013 John Wiley & Sons Ltd
2 mm
738 R. D. STAPL ES ET AL .
0.9
(a)
Fe/(Fe+Mg)
0.8
0.7
(or) (c)
Annulus
0.6
(ir)
Grt 2
0.3
Annulus
Xalm
Grt 1
(ir) (c)(or)
Grt
2
Xpyp
0.2
0.1
0
Xgrs
Xsps
0
1.0
2.0
1.0
3.0
5.0
4.0
(b)
Fe/(Fe+Mg)
0.9
0.8
Xalm
0.7
Grt 2
0.3
Grt 2
Grt 1
Xgrs
0.2
0.1
0
Xpyp
Xsps
0
1.0
2.0
4.0
3.0
5.0
P– T – D H I S T O R Y
(c)
0.9
Stage-1 P–T–d history
Fe/(Fe+Mg)
0.8
0.7
Xalm
0.6
Grt
2
0.3
Grt
2
Grt 1
0.2
Xgrs
0.1
Xpyp
0
retrograde diffusional re-equilibration during resorption of garnet (Kohn & Spear, 2000). Within sample
1, the replacement of Grt-1 and the outer portion of
Grt-2, as well as the interface (annulus) between
them, by the same assemblage (St-Plg-Mnz) suggests
garnet was resorbed both from the inside and outside
following growth of both Grt-1 and Grt-2. This is
supported by the similarity in the composition of the
complexly zoned garnet at its interface with the
replacement products in the core, annulus, and outer
atoll locations–a composition attributed above to
garnet resorption during retrogression. The slightly
higher Fe/(Fe + Mg) and grossular contents within
the core of Grt-1 from sample 1 suggest incomplete
diffusional re-equilibration across Grt-1 during retrogression.
More calcic plagioclase occurs adjacent to resorbed
portions of garnet in all samples and at the rims of
matrix grains, and thus, may likewise reflect the precipitation of new plagioclase upon garnet resorption
during decompression (e.g., Spear et al., 1990). The
alternative that this more calcic plagioclase is due to
the progressive consumption of a Ca-bearing phase
such as epidote is considered less likely as this mineral is not preserved as inclusions in garnet. The less
calcic cores of matrix plagioclase and those at a distance from garnet likely record equilibrium compositions along the prograde path or an earlier
metamorphic event. Thus, the most reliable P–T conditions derive from the composition of minerals associated with the calcic plagioclase. Conventional
barometric calculations will therefore yield a minimum estimate of near-peak P–T conditions.
Xsps
0
1.0
2.0
3.0
mm
Figure 4. Compositional profiles across garnet porphyroblasts.
(a) Sample 1. Abbreviations for Grt-2: ir, inner rim; c, core;
or, outer rim. Note the sharp compositional gradient between
vertical dashed lines. (b) Sample 2. (c) Sample 3. Grey shading
shows the boundaries of garnet 1 and garnet 2.
The earliest phase of deformation and metamorphism
is recorded in the syn-kinematic, inclusion-rich core
(Grt-1) of a two-stage garnet. A sigmoidal inclusion
trail within the garnet core indicates non-coaxial
shear, synchronous with the development of the dominant penetrative foliation (ST). Due to the lack of
suitable inclusions within the Grt-1 core, the preserved growth zoning within samples 2 and 3 is the
only indicator of the P–T evolution during the first
stage of garnet growth. Garnet isopleth thermobarometry (e.g. Vance & Mahar, 1998) was used to estimate the P–T conditions of this earliest garnet
growth event. Isopleths of spessartine, almandine and
pyrope corresponding to the Grt-1 core composition
in sample 2 intersect tightly at ~600 °C and 8 kbar
(Fig. 6). Compositional isopleths corresponding to
the centre of Grt-1 in sample 3 do not intersect at a
point, but define a narrow field from 565 to 585 °C
and from 7.8 to 8.5 kbar. The near vertical slope of
XPyp isopleths implies Grt-1 of samples 2 and 3 grew
along a P–T path of increasing temperature. The out© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 739
l
1.2
5
il P
tS t
tB R
Gr Ilm
1.0
Grt Bt Ky
Pl Ilm
0.8
mol Grt
3
Grt Bt Sil
Pl Ilm
l Ilm
rd P
il C
Bt S
A
B
Grt Bt Crd
Opx Pl Ilm
14
550
650
Temperature (°C)
5
750
Fe/(Fe+Mg)
3
11
11
A
5
2
7
B
B
5
Xgrs
mol St
3
450
A
0
0.00.06 .1
4
7
9
0.0
0.18
9
0.16
1 - Chl Ms Pl Ab Ep Spn
2 - Chl Ms Ab Ep Spn
3 - Chl Bt Ms Ab Ep Spn Ilm
4 - Chl Bt Ms Ab Ep Ilm
5 - Grt Chl Bt Ms Ab Ep Spn Ilm
6 - Grt Chl Bt Ms Ab Ep Ilm
7 - Grt Chl Bt Ms Ab Ep Rt
8 - Grt Chl Bt Ms Pl Ilm Rt
9 - Grt Bt Ms Pl Ilm
10 - Grt Bt St Pl Ilm Rt
11 - Grt Bt Ms Pg Pl Rt
12 - Grt Bt Ms Pl Rt
13 - Chl Bt St Pl Ilm
14 - Grt Chl Bt Crd Pl Ilm
15 - Grt Bt St Crd Pl Ilm
7
0.5
0.4
0.4 6
2
0
0.3 .38
4
13
0.8
Grt Bt Crd
Pl Ilm
0.76
0.88
9
15
0
450
Grt
0.2
3
1
B
Grt Chl
Bt St
Pl Ilm
0.96
5
2
Chl Bt Ms
Pl Ab Ilm
11
Chl Bt Ms
Pl Ilm
0.78
Grt Bt St
Pl Ilm
0.
0.118
4
lB
tM
sP
(3)
B
Grt Bt Ky
Pl Ilm Rt
0.9
0
0.8 .86
0.82 4
pI
10
lE
7
7
0.94
0.92
Chl
Bt
M2
Ch
Chl Ms
Pg Ab
Ep Spn
1.8
A
lm
3
A
1.3
Grt Bt Ky
Pl Rt
2.0
1.4
Ilm
l Ep
Ms
P
9
Grt Chl Bt
Ms Pl Ilm
9
6
(1)
Grt
4
Chl Bt M
s
(2) Ilmh Pl Ab
Pressure (kbar)
11
6
9
0
0.
Grt Bt Ky
Ms Pl Rt
0.2
Chl Bt
Ms Ab
Ep Spn
2.2
1.
12
Grt Bt Ms
Pl Rt
5
11
1.8
8
Grt Chl Bt
Ms Ab Ep
Spn
11
All assemblages + Qtz + H2O
7
4
1.
Sample 1
550
650
3
750 450
550
650
750
Figure 5. Isochemical phase diagram and molar isopleth sections calculated from the normalized effective bulk rock composition
(Table 2b) of sample 1 in the chemical system MnNCKFMASHTO. H2O and quartz calculated as in excess. Dashed lines display
stage-2 garnet multi-equilibria thermobarometric estimates, with the stage-2 peak metamorphic estimate labelled M2, and each
equilibria numbered as referenced in text. Our preferred stage-2 P–T path interpretation is shown by a dashed arrow (see text for
discussion). The molar abundances of the main phases, shown in Fig. 7, are calculated along the path from points A to B.
ward decreasing grossular content within Grt-1 indicates these rocks followed a prograde P–T trajectory
along shallower slope than XGrs isopleths. Due to the
absence of monazite inclusions within the garnet
core, there is presently no upper age constraint on
the timing of this metamorphic and deformational
event. Grt-1 growth may be as old as Late Permian,
similar to the metamorphic domain to the west, but
© 2013 John Wiley & Sons Ltd
we cannot rule out that it may be as young as Late
Jurassic in age.
Stage-2 P–T-d history
Growth of euhedral to subhedral inclusion-poor garnet (Grt-2) during a discrete, later metamorphic event
is supported by the following: (i) a large chemical
11
26
0.7
27
28
M2
0.8
M1
7
Bt And
Crd Pl Ilm
600
Bt Sil Crd
Pl Ilm
650
31
700
Sil Crd Kfs
Pl Ilm
0.84
5
0.86
0.88
0.90
750
Fe/(Fe+Mg)
3
17 - Grt Bt St Ms Pl
18 - Grt Bt St Ms Pl Ilm Rt
19 - Grt Bt St Pl Ilm Rt
20 - Grt Bt St Ky Pl Ilm Rt
21 - Grt Bt St Ky Pl Ilm
22 - Grt Bt Ky Ms Ep Rt
23 - Grt Bt Ky Ms Pg Pl Rt
24 - Grt Bt Ky Ms Pl Ilm Rt
25 - Grt Bt Sil Pl Ilm Rt
26 - Grt Bt St Sil Pl Ilm
27 - Grt Bt St Crd Pl Ilm
28 - Grt Bt St Sil Crd Pl Ilm
29 - Bt St Sil Crd Pl Ilm
30 - Grt Bt And Crd Pl Ilm
31 - Bt Sil Crd Kfs Pl Ilm
11
0.12
30
Temperature (°C)
1 - Chl Ctd Ms Pg Mrg Ilm Rt
2 - Chl Ctd Ms Pg Mrg Ep Rt
3 - Chl Ctd Ms Pg Mrg Ep Ilm
4 - Chl Ctd Ms Pg Ep Rt
5 - Grt Chl Ctd Ms Pg Mrg Ep Ilm
6 - Grt Chl Ctd Ms Pg Ep Ilm Rt
7 - Grt Chl St Ms Pg Mrg Ilm
8 - Grt Chl St Ms Pg Mrg Ilm Rt
9 - Grt Chl St Ms Pg Mrg Ep Rt
10 - Grt Chl St Ms Pl Ilm
11 - Grt Chl St Ms Pg Pl Ilm
12 - Grt Chl St Ms Pg Pl Ilm Rt
13 - Grt Chl St Ms Pg Pl Ep Rt
14 - Grt Chl Bt St Ms Pl Ilm
15 - Grt Chl Bt St Ms Pl Ilm Rt
16 - Grt CHl Bt St Ms Pl Rt
0.78
0.
M2
9
4
0.1
0.1
.1
7
0
2
03
0.0
2
04
550
0.74
8
0.76
9
Grt Bt Sil
Crd Pl Ilm
29
d
An lm
St l I
Bt rd P
C
4
tB
tK
yM
Gr
Grt Bt Sil
Pl Ilm
Grt Bt St
Pl Ilm
0.
mol Grt
0.
XPyr
tS
tM
sP
Ch
l Ilm
lB
tS
tP
l Ilm
lB
3
M1
0.01
6
500
Grt Bt Ky
Pl Ilm
0.0
3
Ch
lC
td
Chl St
Ms Pl
Ilm
0.8
0.4
0.82
Ms
Mr
gP
l Il
m
5
10
25
1.2
5
0.88
0.90
0.92
Chl Ctd
Ms Pg
Mrg Ilm
1
14
XG
rs
2
21
Grt Chl Bt St Pl Ilm
Ctd Ms
Pg
Mrg
Ilm
Ch
7
17
1.6
Grt Bt Ky
Pl Ilm Rt
20
19
7
0.78
15
11
3 Grt Chl
24
18
2.0
0.79
M1
7
Grt Bt Ky
Pl Rt
0.22
0.2
0.18
0.16
lm
5
12
M1
0.94
Pressure (kbar)
4
8
XA
9
(2)
2.4
M2
1.6
1.20.8
(3)
(1)
M2
13 Grt Bt St
16 Ms Pl Rt
9
6
23
2.0
9
22
Grt Bt Ky
Ms Pl Rt
2.4
Chl Ctd
Ms Pg
Ep Rt
Grt Chl St
Ms Pg
Ep Rt
2.8
0.4
Grt Chl Ctd
Ms Pg
Ep Rt
11
0.8
11
All assemblages
+ Qtz + H2O
1.2
Grt Chl
Ky Ms Grt Bt Ky
Pg Ep Ms Pg
Rt Ep Rt
Sample 2
sP
gP
lR
t
740 R. D. STAPL ES ET AL .
5
Xgrs
3
500
600
700
Figure 6. Isochemical phase diagram and molar isopleth sections calculated from the normalized effective bulk rock composition
(Table 2b) of sample 2 in the chemical system MnNCKFMASHTO. H2O calculated as in excess. P–T conditions of the incipient
stages of stage-1 garnet growth (M1) are estimated from the intersection of XAlm (dashed), XPyr (dotted), and XGrs (dotted/dashed)
isopleths corresponding to the Grt-1 core composition of sample 2. Dashed lines display stage-2 garnet multi-equilibria
thermobarometric estimates, with the stage-2 peak metamorphic estimate labelled M2, and each equilibria numbered as referenced
in text. The interpreted stage-2 P–T path is shown by a dashed arrow.
gradient/discontinuity at the boundary between the
inclusion-rich core (Grt-1) and inclusion-poor overgrowth (Grt-2); and (ii) an increase in Fe/(Fe + Mg)
and XGrs, and a decrease in XAlm and XPyp values
from Grt-1 to Grt-2, which does not conform to a
prograde path involving garnet growth (Fig. 6).
Multi-equilibria thermobarometric estimates are
presented in Figs 5 & 6 and Table 3. In order to
minimize the effects of retrograde diffusion, near
thermal peak thermobarometric estimates of stage-2
garnet growth were calculated from minimum Fe/
(Fe + Mg) values just inboard of the Mn- and
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 741
Fe-enriched outer rim, together with the rims of
plagioclase and the least Fe-enriched biotite. Thermobarometric estimates of stage-2 garnet growth for samples 1–3 range between 650–680 °C and 8–10 kbar.
Garnet and plagioclase in samples 1 and 2 have low
grossular and anorthite contents, which reduces the
reliability of the geobarometric estimates (Ashworth &
Evirgen, 1985; Todd, 1998). Part of the inaccuracy
associated with low-anorthite contents may have been
addressed here by using the plagioclase model of
Aranovich (1991), which Aranovich found yields lower
pressures for sodic plagioclase that are more consistent
with independent estimates. Despite a potential inaccuracy associated with low grossular contents for Grt-2
of samples 1 and 2, the barometric estimates are consistent with the more reliable results of the more calcic
mineral compositions in sample 3. Florence & Spear
(1993) noted that diffusional modification of a sharp
compositional gradient within garnet due to an intermittent period of garnet dissolution during staurolite
growth along the prograde path will shift Fe/
(Fe + Mg) at the garnet rim to a higher value, yielding
a calculated temperature lower than the true thermal
peak. Therefore, the slightly higher peak P–T estimates
of 680 °C and 9–10 kbar from sample 3, in which staurolite is absent, may be a closer estimate to near-peak
stage-2 P–T conditions.
Conventional multi-equilibria thermobarometric
estimates of peak stage-2 garnet growth are consistent
with the presence of kyanite in sample 2. However,
the P–T estimates of sample 2 lie on the upper thermal
stability limit of staurolite (Fig. 6), yet well within
overall uncertainties of thermobarometric and phase
diagram calculations. Nevertheless, garnet-biotite
thermometric estimates may overestimate peak temperatures if biotite was enriched in Fe during garnet
resorption. Arguing against this caveat is the abrupt
decrease of Fe/(Fe + Mg) in biotite >300 lm from
garnet (see above), with little Fe/(Fe + Mg) variation
beyond this distance, suggesting that Fe-enrichment of
biotite during garnet resorption was fairly restricted.
Therefore, temperatures calculated from biotite with
the lowest Fe/(Fe + Mg) composition are expected to
yield a close approximation to peak temperatures.
Alternatively, a small apparent discrepancy between
thermobarometric estimates and equilibrium assemblage modelling may reflect the metastable persistence
of staurolite above its upper thermal stability limit
(e.g., Waters & Lovegrove, 2002; Pattison & Tinkham,
2009; Pattison et al., 2011). The syn- to post-kinematic staurolite and kyanite in samples 2 and 4 is
interpreted to have grown and equilibrated with
Grt-2. This is based on the agreement between Grt-2
thermobarometry and the predicted stability fields
of staurolite and kyanite, as well as the syn- to postkinematic nature of staurolite, kyanite and Grt-2.
By contrast, the origin of prograde, matrix staurolite
in sample 1 differs from the stage-2 paragenesis and its
possible metastable existence in sample 2. Contrary to
© 2013 John Wiley & Sons Ltd
the near ubiquitous static recrystallization textures
associated with stage-2 metamorphism, matrix staurolite of sample 1 has an anhedral and skeletal nature,
and is aligned parallel to the foliation and is deflected
around garnet (Fig. 3d), indicative of syn-kinematic
growth. Though it remains unknown as to which garnet generation this staurolite is deflected around, the
syn-kinematic stage-1 garnet core within other samples
seems to be a more likely candidate. Based on these
textures, it is interpreted that matrix staurolite within
sample 1 is from an earlier metamorphic event, likely
related to stage-1 garnet.
Stage-2 retrograde P–T history
The P–T path is constrained for sample 1 by the production of staurolite from garnet and from the Ca-rich
composition (XGrs = 0.25) of the Grt-2 core. Although
it is possible to produce staurolite with isobaric heating
up to the peak temperature (M2, Fig. 5), this path is
interpreted as unlikely since garnet resorption is negligible even near the metamorphic peak (Fig. 5). Preference is given to a clockwise, looping P–T path in
which staurolite growth and garnet resorption accompany the onset of decompression after peak P–T conditions. The retrograde segment of the path is
constrained to be consistent with the absence of retrograde sillimanite in samples 1 (appearing above 620 °C
and 5.4 kbar) and 2, yet consistent with the presence
of cordierite rims around staurolite in samples 2 and 4,
and retrograde sillimanite in sample 4. These observations require decompression below 5 kbar prior to
cooling below 600 °C (Fig. 6).
To highlight the consistency between observed mineral modes and reaction textures with that predicted
for this steep to near isothermal decompression path
within the model system, molar mineral abundances
were calculated for sample 1 (Fig. 7) and plotted
based on the retrograde P–T path between points A
and B in Fig. 5. At 9 kbar and at near- peak conditions of metamorphism (point A in Fig. 5), muscovite
and garnet are stable along with biotite and plagioclase. Early in the retrograde history, at 8.8 kbar,
staurolite growth is predicted, together with plagioclase and biotite, at the expense of muscovite and garnet (Fig. 7). Upon decompressing through 8.6 kbar,
muscovite is entirely consumed, consistent with its
absence in the observed mineral assemblage, and
below which staurolite remains stable. This modelling
is consistent with the observed assemblage of staurolite, plagioclase and Y-rich monazite within resorbed
portions of both stage-1 and 2 garnet of sample 1,
and retrograde biotite recorded texturally as randomly oriented matrix grains and within the cores of
atoll garnet. The presence of monazite within the resorbed garnet, and its otherwise complete absence as
an included phase within both stage-1 and 2 garnet
suggests that this monazite was an accessory product
of these garnet and muscovite-consuming reactions.
742 R. D. STAPL ES ET AL .
A
staurolite grains grew at the expense of garnet following growth of the stage-2 garnet rims.
B
13.0
MONAZITE GEOCHRONOLOGY
Pl
Geochronological methods
Molar abundance
12.0
2.0
Prior to U-Th-Pb analysis, the locations of all monazite grains in thin section were identified using an automated scanning routine on a Zeiss EVO 50 series
scanning electron microscope (SEM) at the Geological
Survey of Canada in Ottawa, operating at 20 kV accelerating voltage and 500 pA beam current. Back scattered electron (BSE) images of the in situ monazite
grains were also obtained to provide insight into their
petrological context and internal zoning, identify
cracks and mineral inclusions, and to guide analytical
spot placement. In order to better characterize chemical zonation and potential age domains within individual monazite grains, chemical X-ray maps of Y, U, Th
and Ca in strategically selected monazite grains were
produced using a Cameca SX50 electron microprobe
at the University of Massachusetts operating at a high
current (240–260 nA), with small step sizes (0.25–
0.62 lm), and rastering of the electron beam. For each
of the Y maps, the approximate Y concentration in
ppm for each pixel was calculated using the AgeMap
program (Williams et al., 1999; modified by Goncalves
et al., 2005) in order to estimate the approximate Y
concentration for each SHRIMP spot (Table 4).
Bt
Grt
1.0
Ms
St
0.0
8
9
7
6
Pressure (kbar)
Figure 7. Diagram showing the evolution of the molar
abundance in sample 1 of the main phases as calculated along
the P–T path from point A to B in Fig. 5.
As discussed below, the consistently younger ages and
distinctly high Y-content of monazite intimately associated with euhedral, inclusion-free staurolite within
resorbed portions of both stage-1 and -2 garnet of
sample 1, corroborates the interpretation that these
Table 4. SHRIMP U-Th-Pb analytical data for monazite.
Spota
Sample 1
M131.1
M102.1
M102.2
M9.1
Sample 2
M55.1
M55.2
M156.1
M157.1
M157.2
M118.1
M118.2
Sample 4
M144.1
M123.1
M107.1
M246.1
M246.2
M511.1
M511.2
M511.3
M516.1
M516.2
M516.3
f(206)207
(%)d
Y
(ppm)c
Th/U
Matrix || ST
Replaced Grt
Replaced Grt
Replaced Grt
core
core
rim
core
400
1400
3400
1700
8.64
7.43
6.39
2.82
1.039E-03
1.144E-03
2.215E-03
4.817E-04
1.143E-04
1.716E-04
1.329E-04
6.263E-05
1.67
1.64
4.01
1.15
0.0594
0.0591
0.0780
0.0552
0.0006
0.0010
0.0026
0.0005
52.316
54.822
56.004
56.010
0.628
0.877
0.672
0.672
120.4
115.0
109.9
113.1
1.4
1.8
1.4
1.4
St
St
St
St
St
Matrix k ST
Matrix k ST
core
rim
rim
N/A
N/A
core
rim
5800
5200
8500
8500
6500
2300
5500
5.35
6.44
7.50
9.19
4.18
4.53
6.71
6.104E-04
8.427E-04
9.815E-04
1.112E-03
4.238E-04
4.112E-04
7.765E-04
6.714E-05
9.269E-05
1.080E-04
1.223E-04
5.085E-05
6.169E-05
9.317E-05
0.98
1.52
1.68
1.45
9.33
0.77
1.43
0.0539
0.0582
0.0595
0.0576
0.0535
0.0522
0.0574
0.0004
0.0006
0.0011
0.0009
0.0004
0.0006
0.0006
53.241
54.034
54.207
53.001
53.004
53.624
55.046
0.639
0.540
0.759
0.742
0.583
0.590
0.661
119.1
116.8
116.2
119.1
119.7
118.5
114.7
1.5
1.2
1.6
1.7
1.3
1.3
1.4
Ky || ST
Matrix || ST
Crd || ST
Crd || ST
Crd || ST
Matrix
Matrix
Matrix
Matrix
Matrix
Matrix
rim
core
rim
core
rim
core
rim
core
rim
core
rim
11000
14000
10 000
15 000
10 000
13 500
3700
13 500
6800
17 300
6800
6.45
2.83
7.71
4.63
5.84
5.73
8.69
5.21
7.50
5.17
9.71
9.508E-04
3.323E-04
8.232E-04
2.726E-04
5.949E-04
4.610E-04
8.827E-04
4.710E-04
6.238E-04
3.895E-04
8.472E-04
9.508E-05
3.988E-05
9.055E-05
6.271E-05
9.518E-05
7.376E-05
8.827E-05
5.181E-05
9.357E-05
5.842E-05
7.625E-05
1.50
0.79
1.10
0.76
1.42
1.06
1.27
0.83
1.05
0.77
1.30
0.0580
0.0523
0.0548
0.0521
0.0574
0.0545
0.0562
0.0527
0.0544
0.0522
0.0564
0.0005
0.0004
0.0010
0.0004
0.0005
0.0004
0.0005
0.0004
0.0011
0.0007
0.0005
53.439
52.814
54.144
55.165
55.265
53.780
54.366
53.599
54.233
53.687
54.064
0.748
0.634
0.653
0.591
0.653
0.743
0.577
0.595
0.875
0.842
0.617
118.1
120.4
117.0
115.3
114.3
117.9
116.3
118.5
116.9
118.4
117.0
1.6
1.4
1.4
1.2
1.3
1.6
1.2
1.3
1.9
1.8
1.3
204
Pb/206Pb
207
Total
Pb/206Pb
Total
U/206Pb
Age (Ma)e,f
206
Pb/238U
Grain location
Textureb
238
Spot: M157.2 = 2nd spot on monazite grain #157.
Texture: location of monazite as inclusion in St, Ky, Crd, or as a matrix grain (|| = elongate monazite grain parallel to foliation).
Y: approximate Y concentration calculated from Y X-ray maps using the AgeMap program as modified by Goncalves et al. (2005).
d
f(206)207 refers to the fraction of total 206Pb that is common Pb, calculated using the 207Pb-method.
e
Ages have been corrected for common Pb using the 207Pb-method.
f
Uncertainties reported at 1r (absolute) and are calculated by numerical propagation of all known sources of error (Stern & Berman, 2000).
a
b
c
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 743
In situ U-Th-Pb analyses using the SHRIMP II at
the Geological Survey of Canada in Ottawa were
performed on monazite cored from polished thin sections and mounted in epoxy together with pre-polished
monazite standards according to the methods of Rayner & Stern (2002). Targeted areas of monazite were
analyzed using a mass-filtered O2 primary beam
focused with a Kohler aperture to a spot measuring
9 9 12 lm. The methods employed follow the
SHRIMP analytical protocols described in detail by
Stern (1997), Stern & Sanborn (1998) and Stern & Berman (2000). Tera-Wasserburg and Concordia plots,
data regression and weighted mean calculations were
made using the program ISOPLOT 3.7 (Ludwig,
2008). Errors assigned to SHRIMP U–Th–Pb ages
were determined using numerical propagation of all
known sources of error as outlined by Stern (1997),
Stern & Sanborn (1998), and Stern & Berman (2000).
Uncertainties for individual analyses (ratios and ages)
shown in Tables 4 and S1, and error ellipses shown in
Fig. 8, are presented at the 1r level, whereas weighted
mean ages and 2r uncertainties are provided in the
text, Table 3 and Fig. 8.
(a)
Sample 1
Geochronological results
A plot of the monazite SHRIMP data for sample 4
on a conventional (Wetherill) U-Pb concordia diagram reveals a reverse discordance exhibited by some
monazite analyses (Fig. 8d). Excess 206Pb due to
incorporation of 230Th into monazite at the time of
crystallization will lead to an overestimation of
206
Pb/238U ages and reverse discordance paralleling
the 206Pb/238U axis on conventional U-Pb concordia
diagrams (Sch€arer, 1984). Our data show a subtle
trend of increasing reverse discordance with increasing 208Pb/206Pb (the radiogenic proxy of Th/U),
which indicates that to some minor degree the reverse
discordance may be the result of excess 206Pb due to
230
Th disequilibrium. Unfortunately the 208Pb/232Th
chronometer, which is ideal for monazite since it is
not known to be affected by isotopic disequilibrium
(i.e., unsupported 206Pb), cannot be used in this study
because one of the three monazite standards routinely yielded a high Th-Pb age possibly due to an
indeterminate matrix effect. This creates an unquantifiable uncertainty in the Th-Pb ages for the monazite.
(b)
Regression intercept at
112.4 ± 1.7 Ma
0.08
Sample 2
Regression intercept at
117.8 ± 1.0 Ma
0.062
MSWD = 2.8
MSWD = 1.7
108
0.054
207Pb-corrected ages
0.05
122
118
114
MSWD = 1.8
POF = 0.100
117
115
113
207Pb-corrected ages
110
126
0.04
50
54
52
56
58
0.046
50
60
122
52
238
U/206Pb
(c)
Sample 4
238
(d)
Regression intercept at
117.2 ± 0.9 Ma
118
114
54
56
Sample 4
130
123
206
207
117.1 ± 1.2
MSWD = 1.4
POF = 0.15
119
117
115
113
207Pb-corrected ages
0.050
126
122
118
238
Pb/ U
238
0.054
120
Weighted mean age
121
0.018
110
206
06
Pb/ U age (Ma)
0.058
58
U/206Pb
0.020
MSWD = 1.4
Pb/2 Pb
117.8 ± 1.8
0.050
126
0.062
Weighted mean age
121
119
238
112
110
Pb/ U age (Ma)
114
0.058
206
MSWD = 2.9
POF = 0.057
206
206
112.3 ± 6.2
Pb/ Pb
238
0.06
Weighted mean age
116
207
Pb/ U age (Ma)
0.07
207
Pb/206Pb
118
0.016
100
114
90
0.046
50
52
54
238
206
U/ Pb
56
58
0.014
0.09
0.10
0.11
0.12
0.13
0.14
207
Pb/235U
Figure 8. U-Pb isotopic plots for samples 1, 2 and 4. (a–c) Tera–Wasserburg plots for samples 1, 2 and 4 showing isotopic data
uncorrected for common Pb. Regression is fitted through a Late Cretaceous common Pb isotopic composition (Stacey & Kramers,
1975). Dashed ellipse excluded from the regression (see text for discussion). Error ellipses represent 1r level of uncertainty. Inset
shows distribution of (207Pb-corrected) 206Pb/238U ages, with error bars at 1r. Bold horizontal line is reference mean age. (d)
Concordia plot for monazite from sample 4. Note the horizontal trend parallel to the 207Pb/235U axis in the discordant data.
© 2013 John Wiley & Sons Ltd
744 R. D. STAPL ES ET AL .
However, this problem seems to have affected only
one-third of the standards, so it is likely some, or
possibly all, of the monazite unknowns were similarly
unaffected. Consequently, given that the 206Pb/238U
ages are consistently younger than 208Pb/232Th ages
(Table S1) would suggest these grains do not contain
significant excess 206Pb. Additionally, when applying
arer (1984) it at
the excess 206Pb correction of Sch€
most produces an excess in age of 1.2–1.9 Ma, which
is no greater than analytical error, even for monazite
that grew from a metamorphic fluid with extremely
low Th/U (0.3–0.7) and had a large amount of Th
fractionation. Thus, despite any minor excess of
206
Pb, we are confident the 206Pb/238U ages will provide reasonably accurate constraints within the resolution required for this study to differentiate the
metamorphic and deformation events.
We can infer a further reduction in the significance
of the contribution of excess 206Pb to reverse discordance based on analyses from sample 4. These analyses do not plot in a vertical trend above concordia,
as would be expected if the reverse discordance was
due to excess 206Pb. Rather, the analyses trend to the
left of concordia parallel to the 207Pb/235U axis
(Fig. 8d). This trend is interpreted to be largely the
result of an overcorrected 207Pb/235U ratio using the
204
Pb method. The errors arising from low 204Pb
counts, background interference and a 204 isobar,
can overcorrect for common Pb, which most
adversely affects young samples with relatively low
concentrations of 207Pb.
For the sake of thoroughness, the analyses were
corrected for common Pb based on the 204Pb (Table
S1) and 207Pb (Table 4) methods following the procedure of Stern & Berman (2000) and Ireland & Gibson (1998). The two correction methods yield ages
indistinguishable within error. However, all things
considered, the 206Pb/238U chronometer corrected
using the 207Pb method is thought to provide the
most meaningful ages for this study; accordingly, all
ages quoted and displayed on Tera-Wasserburg concordia diagrams (Fig. 8a–c) are based on the
206
Pb/238U chronometer.
matrix monazite yields an age of 120.4 1.4 Ma.
Grain 102 (Fig. 9a) is a subhedral, 20 9 45 lm grain
that occurs with plagioclase and euhedral staurolite
within a resorption ring between Grt-1 and -2. Grain
102 has a uniform, moderate-Y (700–1500 ppm) core
and high-Y (2000–3400 ppm) rim, with a sharp
boundary between the two zones. Spot analyses of
the moderate-Y core and high-Y rim yield ages of
115.0 1.8 Ma and 109.9 1.4 Ma, respectively.
Grain 9 (Fig. 9b) is a subhedral, 30 9 50 lm grain
that occurs together with plagioclase and euhedral
staurolite within a resorbed hole inside Grt-1. Monazite nine has a fairly uniform moderate-Y concentration (1000–2000 ppm), and a very thin (<5 lm) highY (3500–4000 ppm) rim. A single analysis from the
centre of grain nine yields a 206Pb/238U age of
113.1 1.4 Ma.
Sample 2
Analyses of monazite within sample 2 were obtained
on three inclusions within two different staurolite
porphyroblasts (Fig. 9d,e), and a single matrix grain.
Two of the monazite inclusions within staurolite are
anhedral to subhedral grains that are ~30 9 100 lm.
The third inclusion within staurolite is very elongated, ~160 lm in the longest dimension oriented
roughly parallel to a weakly defined and statically
overprinted external foliation (ST). Only one of the
three inclusions within staurolite is intersected by a
very fine micro-crack (Fig. 9e). X-ray mapping
revealed three distinct compositional domains within
monazite included in staurolite and in the matrix: (i)
moderate-Y, high-Th core; (ii) an intermediate lowY, low-Th zone; and (iii) high-Y, moderate-Th rim.
Despite the different textural positions and discrete
chemical domains, both the core and outer high-Y
rim yield a restricted range of ages between
119.7 1.3 Ma and 114.7 1.4 Ma (see Table 4)
that are interpreted to form a single population with
a weighted mean age of 117.8 1.7 (MSWD = 1.8)
(Fig. 8b).
Sample 4
Sample 1
Analyses of monazite within sample 1 were obtained
on a single matrix grain, as well as two monazite
crystals that occur together with plagioclase and euhedral staurolite within replacement textures inside
garnet (Fig. 9a,b). Elongated matrix monazite is consistently aligned parallel to ST. The matrix grain,
monazite 131 (Fig. 9c), is acicular, measuring <35 lm
in shortest dimension and up to 165 lm in the longest dimension, and is approximately parallel to the
foliation (ST). Monazite 131 has a broad, uniform
low Y (<700 ppm) core surrounded by a thin
(<5 lm) Y-depleted zone, and a thin (<5 lm) Y-rich
rim. A single spot analysis of the low-Y core of this
Analyses of monazite from sample 4 were obtained
on one inclusion in kyanite, two inclusions in cordierite and two matrix grains. Monazite inclusions within
cordierite and syn-kinematic kyanite are elongate, euhedral to subhedral grains, >30 lm in the shortest
dimension and up to 200 lm in the longest dimension. Elongate monazite inclusions within kyanite and
cordierite, as well as matrix grains, are aligned parallel to ST, which is defined by the rutile trains and
aligned kyanite and staurolite (Fig. 9f). All monazite
grains, regardless of textural location, show similar
concentrations and compositional zoning patterns
characterized by a Y-rich (12 000–18 000 ppm) core
surrounded by a uniform zone relatively depleted in
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 745
St
(a)
Pl
M9
(b)
Grt-1
St
Pl
Pl
10 µm
St
Pl
M102
1
St
Grt-1
Grt
-2
M102
Pl
St
St
2
10 µm
M9
100 µm
100 µm
M131
(c)
M156
(d)
1
20 µm
M157
2
M156
20 µm
ST
St
1
M157
1 mm
20 µm
(e)
M55
200 µm
M246
1
1
50 µm
2
20 µm
2
Crd
M55
(f)
Bt
Ky
St
St
ST
St
Rt
50 µm
(g)
M516
(h)
1 mm
M123
3
1
2
20 µm
1
20 µm
(j)
(i)
2
1
M511
3
20 µm
1
M144
20 µm
Figure 9. SEM backscattered images of monazite and the surrounding area in samples 1, 2 and 4. Ellipses on inset Yttrium (Y) Xray maps of monazite show the location of SHRIMP analysis spots. Lighter shades of grey on Y maps indicate relatively higher
concentration. (a) M102 and (b) M9, both associated with staurolite and plagioclase within heavily resorbed garnet porphyroblasts
in sample 1. (c) Elongate matrix monazite M131 aligned near parallel to the transposition foliation (ST). (d) M156 and M157, and
(e) M55, within staurolite from sample 2. (f) Elongate monazite, M246, aligned parallel to the relict transposition foliation (ST) in
sample 4, which is defined by rutile trains within the matrix, and as inclusion trains within porphyroblasts of kyanite, staurolite
and cordierite. (g–j) Y maps of monazite grains from sample 4.
© 2013 John Wiley & Sons Ltd
746 R. D. STAPL ES ET AL .
Y (6000–10 000 ppm) (Fig. 9g–j). This Y zoning is
mimicked by Th. In spite of the observed textural and
chemical differences described here, all 11 analyses of
these five monazite grains form a single population
with a weighted mean 206Pb/238U age of 117.1 1.2 Ma
(MSWD = 1.4) (Fig. 8c).
In contrast to samples 1 and 2, there is no high-Y
rim observed in monazite from sample 4. This is
interpreted to be due to the relatively small amount
of garnet observed in sample 4. In this garnet-poor
sample there would not be a significant amount of Y
tied up in garnet, nor excess Y released upon garnet
breakdown.
Interpretation of geochronological data – constraining a
Cretaceous P–T path
It is unlikely that the monazite ages have been reset
by diffusive Pb loss. Based on experimentally determined diffusion parameters, Cherniak et al. (2004)
and Gardes et al. (2006) have determined that monazite would have to be exposed to temperatures
>800 °C for a geologically unrealistic length of time
for there to be any appreciable Pb loss, and predict
closure temperatures comparable to zircon (i.e.
900 °C). These conclusions are consistent with
numerous observations of sharp boundaries between
domains of contrasting Pb concentrations, and the
preservation of significantly older ages through younger granulite facies metamorphic events (e.g. DeWolf
et al., 1993; Spear & Parrish, 1996; Braun et al.,
1998; Cocherie et al., 1998; Crowley & Ghent, 1999;
Zhu & O’Nions, 1999). Given the peak temperature
estimates for these samples (~650–680 °C), we interpret that these grains have not been reset due to Pb
loss by thermally-activated volume diffusion.
More plausibly, several studies (Seydoux-Guillaume et al., 2002; Harlov & Hetherington, 2010;
Hetherington et al., 2010; Harlov et al., 2011; Williams et al., 2011) have shown that monazite ages
may be reset by fluid-assisted coupled dissolutionreprecipitation, which may operate at temperatures
well below the closure temperature of monazite.
Unlike the matrix grains, monazite inclusions in staurolite are shielded from an intergranular fluid, and
should therefore retain their primary metamorphic
age. However, if monazite inclusions in staurolite are
intersected by a microcrack within the third dimension, this could provide a pathway for fluid assisted
dissolution, reprecipitation and Pb loss. Williams
et al. (2011) examined the resetting of monazite ages
during fluid-related coupled dissolution-reprecipitation, and observed that altered rims yield a scattering
of reset ages, whereas the age of unaltered monazite
cores were largely unaffected. Contrary to the results
of Williams et al. (2011), the ages obtained from distinct core-rim compositional domains of monazite
inclusions in staurolite from this study (Fig. 9e) agree
within error (Table 4), leaving no indication that
these grains have been affected by fluid-assisted coupled dissolution-reprecipitation. Ages from these
grains are therefore interpreted as primary metamorphic ages. Furthermore, because monazite both in
the matrix and included in staurolite yield a single
age population, we interpret matrix monazite to also
preserve their primary metamorphic age.
Timing of stage-2 metamorphism and deformation
Nineteen 206Pb/238U analyses from all 11 monazite
grains (excluding the Y-rich monazite within resorbed
Grt-2 of sample 1) yield a single age population at
117.5 0.9 Ma (MSWD = 1.7). The incorporation
of c. 118 Ma monazite within syn- to post-kinematic,
stage-2 porphyroblasts of staurolite and kyanite indicates this population of monazite grew prior to, or
synchronous with, these near-peak stage-2 porphyroblasts. As discussed below, slightly younger ages (c.
112 Ma) from distinct, Y-rich monazite within resorbed portions of both stage-1 and 2 garnet of sample 1, are interpreted to have formed as garnet broke
down during the onset of decompression following
the peak of stage-2 metamorphism. The absence of
monazite included in stage-1 garnet, and the limited
time between prograde metamorphism (c. 118 Ma)
and stage-2 decompression (c. 112 Ma) suggest it is
unlikely the c. 118 Ma monazite population grew
during stage-1 metamorphism. Instead the c. 118 Ma
monazite population is interpreted to date stage-2
prograde metamorphism. Furthermore, the common
alignment of monazite parallel to ST in the matrix
(Fig. 9c,f), and its inclusion in syn- to post-ST porphyroblasts of staurolite and kyanite, suggests monazite grew during the waning phase of protracted
development of ST during prograde stage-2 metamorphism. The partially syn-kinematic nature of both
Grt-1 and stage-2 staurolite and kyanite porphyroblasts with respect to ST, suggests ST is a composite
D1/D2 transposition foliation in the sense of Williams
(1983) and Tobisch & Paterson (1988). The presence
of intrafolial isoclinal folds suggests this composite
transposition foliation developed by isoclinal folding
and rigid body rotation of the older D1 foliation into
parallelism with the new D2 foliation, such that the
elements defining the new foliation (ST) are inherited
from the old foliation (cf. Williams, 1983).
Approximately 1.2 km to the east of sample 2,
slightly older Early Cretaceous dates were obtained
by Berman et al. (2007a), who calculated an average
206
Pb/238U age of 146 3 Ma from three different
monazite grains (sample 4 of Berman et al., 2007a).
Berman et al. interpreted these grains as primary
inclusions within post-kinematic staurolite porphyroblasts that grew during a high-P (~9 kbar) metamorphic event. The c. 117.5 0.9 Ma age determined in
this study records the time at which these samples
passed through the monazite growth reaction, which
is a function of pressure, temperature and the bulk
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 747
composition. Given the considerable chemical variability between sample 4 of Berman et al. (2007a)
(Ca/Na = 0.14) and that of samples 2 (Ca/Na = 0.56)
and 4 (Ca/Na = 0.99) of this study, there is no reason
to expect that the monazite-producing reactions were
the same between the two studies. Nor should we
expect that the reactions were operating under the
same P–T conditions, and hence time. For example,
Fitzsimons et al. (2005) demonstrated that variations
in the evolving metamorphic mineral assemblages,
between samples of contrasting composition in AFM
space, resulted in the growth and preservation of
monazite from different metamorphic reactions and
under considerably different P–T conditions (e.g.
greenschist- v. amphibolite facies). Janots et al.
(2008) also found that monazite stability is sensitive
to bulk composition. In their study of the stability
and phase relations between allanite and monazite,
Janots et al. (2008) offered the following generalized
prograde reaction: Allanite + apatite + Al-Fe-Mg
phases1 = monazite + anorthite + Al-Fe-Mg phases 2.
Janots et al. (2008) suggested that metapelites with
a high Ca/Na ratio, and subsequently high anorthite
activity in plagioclase, would shift the above equilibrium to the right, stabilizing allanite to higher temperature. Given the considerably higher Ca/Na and
XAn contents from samples of the present study
v. that of sample 4 of Berman et al. (2007a), it follows
from Janots et al. (2008) that the monazite-after-allanite equilibrium would occur at higher temperatures in
rocks of this study. Consequently, monazite in our
study should yield younger ages for rocks following a
clockwise P–T path, precisely as observed. We therefore interpret that prograde metamorphism may have
occurred between 146 and 118 Ma in this region. Such
protracted metamorphism, ~28 Ma does not seem
unreasonable considering recent work which document
a 25 Ma duration of eclogite facies conditions (Mattison et al., 2006), and prograde garnet growth sustained over a 25–40 Ma interval (Skora et al., 2009;
Cheng et al., 2011).
Timing of stage-2 retrogression and exhumation
Monazite grains M9 and M102 in sample 1 occur
within resorbed portions of garnet associated with
post-kinematic staurolite, and yield a weighted mean
age of 112.3 6.2 Ma (MSWD = 2.9) (Fig. 8a). This
age is nominally younger than the 120.4 1.4 Ma
matrix monazite, but the difference is considered significant given their distinctly higher yttrium content
(~1000–10 000 ppm Y) compared to the older matrix
monazite (<500 ppm). Bea & Montero (1999) and
Pyle & Spear (1999) determined that garnet and xenotime are the only significant reservoirs of yttrium
in metapelitic rocks, and that the growth of garnet is
accompanied by the consumption of xenotime. Given
the large volume of garnet in these rocks, it follows
that the majority of yttrium resides within garnet.
© 2013 John Wiley & Sons Ltd
Therefore, we interpret that the high-Y, younger
monazite formed during the release of yttrium
accompanying breakdown of garnet (e.g. Pyle &
Spear, 1999; Foster et al., 2002), which is interpreted
to have occurred during decompression from the
stage-2 metamorphic peak (~650–680 °C and 9 kbar)
into the staurolite stability field. This conclusion is
supported by the occurrence of 114.3 1.3 Ma monazite within cordierite, which is constrained by phase
relationships to have crystallized below ~5 kbar
(Fig. 6) after c. 114 Ma. It is also consistent with
results from 7–12 km northwest of sample 1, where
115–107 Ma relatively Y-rich monazite in two rocks
formed during garnet resorption and re-equilibration
at pressures less than 6 kbar (Berman et al., 2007a).
Finally, we note that these younger ages directly
linked to decompression from peak stage-2 metamorphism further support our interpretation that the
older ages (c. 118 Ma) for monazite inclusions within
staurolite and kyanite date the prograde history of
the ~9 kbar stage-2 metamorphic event.
TE C T O N I C I M P L I C A T I O N S
Previous work in the Yukon-Tanana terrane of easternmost Alaska and west-central Yukon suggests the
terrane was strongly transposed and metamorphosed
to amphibolite facies in the Permo-Triassic (DuselBacon et al., 1995, 2002; Berman et al., 2007a; Beranek & Mortensen, 2011), followed by a metamorphic
overprint in the Early Jurassic before being exhumed
to upper crustal levels in the Early to Middle Jurassic
(Dusel-Bacon et al., 1995, 2002; Hansen & DuselBacon, 1998; Berman et al., 2007a). However, the
results of this study, together with the work of Berman et al. (2007a), indicate a more protracted and
heterogeneous tectono-thermal history for the
Yukon-Tanana terrane. In particular, rocks in this
part of west-central Yukon appear to be stratigraphically correlative with, and share a similar structural
style, deformation sequence, and metamorphic grade
as the surrounding Permian to Early Jurassic metamorphic domain. However, the Australia Mountain
domain records a transposition event associated with
burial and metamorphism to 8–10 kbar and 650–
680 °C in the Early to mid-Cretaceous (c. 146–
118 Ma), which is not recorded in the surrounding
Permian to Early Jurassic metamorphic domain.
These observations require that the Australia Mountain domain occupied a deeper crustal level in the
Early to mid-Cretaceous than the surrounding Permian to Early Jurassic metamorphic domain, and
therefore provides a window into the Early to midCretaceous infrastructure of the orogen in westcentral Yukon. We apply the term ‘infrastructure’ in
a similar sense to that suggested by De Sitter &
Zwart (1960) and Culshaw et al. (2006) when describing mid- to lower-crustal levels in an orogen characterized by high-grade, shallowly dipping, ductily
748 R. D. STAPL ES ET AL .
SE
NE
C
Willow
Unde
Miss formed &
issip
pian unmetam
Reid
Lake orphosed
s bath
olith
SW
NW
A
Lake fa
ult
Stewart Ri
Austra
ver fault
Early
meta to mid-C
re
morp
hic d taceous
oma
in
B
lia Cre
ek fau
lt
Australia Mtn.
area
deformed and transposed rocks. Conversely, the overlying ‘superstructure’ would be characterized by
upright, brittle structures and low-metamorphic grade.
For us, this rheological contrast between upper and
lower crustal levels is time specific, in our case, the
Early to mid-Cretaceous. We make this distinction
because, as explained above, rocks formerly situated in
the lower crust, the infrastructure, were progressively
exhumed in the Jurassic and were incorporated into
the superstructure above the Early to mid-Cretaceous
infrastructure.
Our study indicates a pattern of structurally downward younging deformation and metamorphism in
the Yukon-Tanana terrane, similar to what has been
described in the southeastern Cordillera (Carr, 1991;
Parrish, 1995; Crowley et al., 2000; Gibson et al.,
2005). There, rocks buried, heated and exhumed in
the Jurassic (Archibald et al., 1983; Brown et al.,
1992; Colpron et al., 1996; Gibson et al., 2005), are
now juxtaposed against structurally deeper rocks that
were progressively buried and heated from Cretaceous to earliest Eocene (Brown & Carr, 1990; Carr,
1991; Parrish, 1995; Gibson et al., 1999, 2005; Crowley & Parrish, 1999; Crowley et al., 2000). Parrish
(1995), Brown (2004) and Gibson et al. (2008) have
all attributed this downward younging tectonism in
the southeastern Canadian Cordillera to the progressive incorporation and burial of material as the burgeoning orogen propagated northeastward towards
the foreland.
Monazite from this study dated at c. 112 Ma records
the terminus of metamorphism and onset of near isothermal decompression and exhumation of deep seated
metamorphic rocks of the Australia Mountain
domain. Exhumation of this domain in the mid-Cretaceous is substantiated by the presence of metamorphic
detritus (quartz with undulatory extinction, and lesser
amounts of muscovite, feldspar and foliated lithic fragments) in the mid-Cretaceous (Albian) Indian River
Formation (Lowey & Hills, 1988), ~35 km to the west.
An isothermal decompression path alone may not be
sufficient to distinguish between normal faulting and
erosion as the dominant mechanism of exhumation
ssic
Jura
Early domain
hic
ian &
Perm etamorp
m
Figure 10. Schematic block diagram
depicting the Cretaceous metamorphic
domain (core complex?) at Australia
Mountain, juxtaposed against the
Mississipian Reid Lake plutonic complex
and the Permian and Early Jurassic
metamorphic domain by the Stewart River
and Australia Creek faults. View is to the
south. Location of cross-section lines
indicated on Fig. 2. Fill patterns are in
legend of Fig. 2.
(Ring et al., 1999). However, the abrupt juxtaposition
of the Cretaceous metamorphic domain at Australia
Mountain against rocks with Early to Middle Jurassic
cooling ages (Hunt & Roddick, 1992) strongly suggests
the Australia Mountain domain is bound on its west
side by a mid-Cretaceous normal fault, herein named
the Australia Creek fault (Figs 2 & 10). Likewise, the
juxtaposition of the Australia Mountain domain to the
southeast against largely undeformed and unmetamorphosed rocks of the Mississippian Reid Lakes complex
is interpreted to have been accommodated along a normal fault, herein named the Stewart River fault (Figs 2
& 10). Both of these faults are interpreted to have
accommodated a significant amount of extensional
exhumation in the mid-Cretaceous. Unfortunately,
poor exposure in the area of the proposed faults has
prevented the identification of corroborating deformation fabrics. Rather, the position of the fault is constrained by the jump in thermochronometric ages, the
presence of a distinct discontinuity in aeromagnetic
data for this region (Hayward et al., 2012), and differences in lithologies (Permian Klondike schist present
on west side of fault, abundant pre-late Devonian marble unit east of the fault).
Some 100 km to the west of Australia Mountain,
rocks in east-central Alaska also record a shift to
extensional tectonics in the mid-Cretaceous. The
majority of metamorphic rocks in east-central Alaska
are parautochthonous North American continental
margin rocks with top-down-to-the-southeast shear
fabrics and mid-Cretaceous cooling ages (Pavlis
et al., 1993; Hansen & Dusel-Bacon, 1998; DuselBacon et al., 2002). These amphibolite facies rocks
with mid-Cretaceous 40Ar/39Ar and K-Ar cooling
ages are juxtaposed across mylonitic shear zones
against greenschist to amphibolite facies allochthonous rocks of Yukon-Tanana terrane with Early
Jurassic cooling ages. Hansen & Dusel-Bacon (1998)
and Dusel-Bacon et al. (2002) interpreted the topdown-to-the-southeast mylonitic shear zones to have
formed during exhumation of the parautochthonous
rocks from beneath the structurally overlying allochthonous Yukon-Tanana terrane.
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 749
The Cretaceous metamorphic domain in the Australia Mountain area likewise represents a window
through a Cretaceous superstructure, with a Palaeozoic to Jurassic metamorphic and deformational history,
revealing a much younger amphibolite facies Cretaceous infrastructure. Late Devonian (c. 363 Ma, Mortensen, 1990) and Early Mississippian (c. 348 Ma,
Ruks et al., 2006) U-Pb ages from the Mount Burnham orthogneiss within the Australia Mountain
domain are consistent with igneous activity known in
both parautochthonous North American margin rocks
and allochthonous Yukon-Tanana terrane in east-central Alaska as described by Dusel-Bacon et al. (2006).
However, if the Australia Mountain domain is of
parautochthonous North American continental margin affinity, then the Australia Creek fault represents a
major terrane bounding extensional fault analogous to
coeval mid-Cretaceous extensional faults described by
Hansen & Dusel-Bacon (1998) and Dusel-Bacon et al.
(2002) in east-central Alaska. Exhumation of structurally deep parautochthonous continental rocks from
beneath the structurally overlying Yukon-Tanana
terrane is highly conceivable considering that 25–30 km
of crustal section (equivalent to the 9 kbar stage-2 peak
metamorphic estimate) was removed in the middle
Cretaceous following an earlier episode of exhumation
of the Yukon-Tanana terrane in the Early Jurassic.
The Cretaceous domains at Australia Mountain
and in east-central Alaska may be akin to the extensional core complexes identified throughout the
North American Cordillera (Coney, 1980; Armstrong, 1982; Coney & Harms, 1984; Parrish et al.,
1988; Struik, 1993), both in style and geological process. However, Cretaceous extensional exhumation in
the northern Cordillera is distinctly older than the
Eocene extension recorded in core complexes in the
southern Canadian Cordillera and northwestern US
(Parrish et al., 1988), and the Oligocene-Miocene
extension in southwestern US (Coney, 1980).
CONCLUSIONS
Garnet growth zoning patterns within compositionally
distinct inclusion-rich core and inclusion-poor rim
domains, separated by an abrupt chemical discontinuity, are interpreted with the aid of modelled compositional and molar isopleths to record a two-stage garnet
growth history, each reflective of a distinct metamorphic event. The early stages of growth of the inclusionrich, stage-1 core (Grt-1) is interpreted to have initiated
at conditions of ~600 °C, 8 kbar along a clockwise P–T
path synchronous with the development of the external
composite transposition foliation (ST). Due to an
absence of monazite inclusions within stage-1 garnet
cores, this event remains undated. Growth of the
relatively grossular-rich, pyrope-poor stage-2 overgrowth (Grt-2) is interpreted to have initiated at higher
pressure, and/or lower temperature, than Grt-1, culminating at near-thermal peak P–T conditions of ~650–
© 2013 John Wiley & Sons Ltd
680 °C and 9 kbar. These peak thermobarometric
estimates calculated from the composition of the rim
of post-kinematic stage-2 garnet overgrowths (Grt-2)
are consistent with the presence of syn- to postkinematic kyanite, and require only minimal reaction
overstepping (<20 °C) to explain the presence of synto post-kinematic staurolite. In situ U-Th-Pb SHRIMP
dating of elongate monazite aligned with ST in the
matrix, and monazite included in syn- to post-kinematic, stage-2 porphyroblasts of staurolite and kyanite
yield a single age population c. 117.5 0.9 Ma. This
age is interpreted to date the waning development of a
reworked composite transposition foliation during
stage-2 prograde metamorphism. These data, together
with the data from Berman et al. (2007a), record a c.
146–118 Ma, 9 kbar amphibolite facies metamorphic
and deformational event in the Australia Mountain
area of west-central Yukon. However, contrary to the
interpretation of Berman et al. (2007a), the younger
mid-Cretaceous monazite ages (c. 112 Ma, this study;
114–107 Ma, Berman et al., 2007a) are not interpreted
to date a distinct low-P contact metamorphic event.
Rather, the presence of consistently young, Y-rich
monazite, intimately associated with retrograde staurolite and plagioclase within resorbed portions of
garnet, is interpreted to date the timing of garnet
breakdown during decompression from the peak of
metamorphism following the stage-2 garnet growth
event. The results presented above indicate that in contrast to the majority of the Yukon-Tanana terrane,
which was deformed and metamorphosed in the
Permo-Triassic and exhumed to upper crustal levels in
the Jurassic, the Australia Mountain domain occupied
a deep crustal level (~25–30 km) in the Early to midCretaceous. This area therefore represents a tectonic
window into Early to mid Cretaceous infrastructure of
the Yukon-Tanana terrane, potentially comparable to
parautochthonous North American continental margin rocks beneath the Yukon-Tanana terrane in eastcentral Alaska, and may be akin to, but older than,
extensional core complexes identified throughout the
North American Cordillera.
ACKNOWLEDGEMENTS
This work was funded by NSERC grants to MC and
DG, and NSERC scholarship to RS. Additional support was provided through the Geological Survey of
Canada’s (GSC) Geomapping for Energy and Minerals Program (GSC contribution # 20130035). We
thank: P. Hunt at the GSC Ottawa for SEM imaging; M. Raudsepp and E. Czech at the University of
British Columbia for assistance with microprobe
analyses; and M. Jercinovic and the UMass microprobe lab for assistance with X-ray mapping of monazite. D. Tinkham is thanked for providing the
thermodynamic data set tcds55 formatted for use
with Theriak-Domino. N. Rayner, B. Davis and T.
Pestaj at the Geological Survey of Canada SHRIMP
750 R. D. STAPL ES ET AL .
lab are thanked for assistance in the acquisition and
reduction of the U-Th-Pb SHRIMP data. The manuscript benefited from discussions with D. Murphy,
J. Mortensen and E. Knight. Constructive reviews by
J. Mezger and C. Dusel-Bacon, and editorial handling by D. Robinson are gratefully acknowledged.
REFERENCES
Aranovich, L.Ya, 1991. Mineral Equilibria of Multicomponent
Solid Solutions. Nauka Press, Moscow.
Archibald, D.A., Glover, J.K., Price, R.A., Farrar, E. & Carmichael, D.M., 1983. Geochronology and tectonic implications of magmatism and metamorphism, southern Kootenay
Arc and neighbouring regions, southeastern British Columbia. Part I: Jurassic to mid-Cretaceous. Canadian Journal of
Earth Sciences, 20, 1891–1913.
Armstrong, R.L., 1982. Cordilleran metamorphic core complexes–from Arizona to southern Canada. Annual Review of
Earth and Planetary Science Letters, 10, 129–154.
Ashworth, J.R. & Evirgen, M.M., 1985. Plagioclase relations
in pelites, central Menderes Massif, Turkey. II. Perturbation
of garnet-plagioclase geobarometers. Journal of Metamorphic
Geology, 3, 219–229.
Bea, F. & Montero, P., 1999. Behavior of accessory phases
and redistribution of Zr, REE, Y, Th, and U during metamorphism and partial melting of metapelites in the lower
crust: an example from the Kinzigite Formation of Ivrea–
Verbano, NW Italy. Geochimica et Cosmochimica Acta, 63,
1113–1153.
Beranek, L.P. & Mortensen, J.K., 2011. The timing and provenance record of the Late Permian Klondike orogeny in
northwestern Canada and arc-continent collision along western North America. Tectonics, 30: TC5017, doi:10.1029/
2010TC002849.
Berman, R.G., 1991. Thermobarometry using multi-equilibrium calculations: a new technique, with petrological applications. Canadian Mineralogist, 29, 833–855.
Berman, R.G. 2007. winTWQ (version 2.3): A Microsoft Windows-compatible software package for performing internally-consistent thermobarometric calculations. Geological
Survey of Canada Open File, 5462.
Berman, R., Sanborn-Barrie, M., Stern, R. & Carson, C.,
2005. Tectonometamorphism at c. 2.35 and 1.85 Ga in the
Rae domain, western Churchill Province, Nunavut, Canada:
insights from structural, metamorphic and in situ geochronological analysis of the southwestern Committee Bay belt.
Canadian Mineralogist, 43, 409–442.
Berman, R.G., Ryan, J.J., Gordey, S.P. & Villeneuve, M.,
2007a. Permian to Cretaceous polymetamorphic evolution of
the Stewart River region, Yukon-Tanana terrane, Yukon,
Canada: P-T evolution linked with in-situ SHRIMP monazite geochronology. Journal of Metamorphic Geology, 25,
803–827.
Berman, R.G., Aranovich, L.Y., Rancourt, D.G. & Mercier,
P.H.J., 2007b. Reversed phase equilibrium constraints on the
stability of Mg-Fe-Al biotite. American Mineralogist, 92,
139–150.
Berman, R.G., Sanborn-Barrie, M., Rayner, N., Carson, C.,
Sandeman, H.A. & Skulski, T., 2010. Petrological and in situ
SHRIMP geochronological constraints on the tectonometamorphic evolution of the Committee Bay belt, Rae Province,
Nunavut. Precambrian Research, 181, 1–20.
Berman, R.G., Rayner, N., Sanborn-Barrie, M. & Whalen, J.
(in press). The tectonometamorphic evolution of Southampton Island, Nunavut: insight from petrologic modeling and
in situ SHRIMP geochronology of multiple episodes of monazite growth. Precambrian Research.
Braun, I., Montel, J.M. & Nicollet, C., 1998. Electron microprobe dating of monazites from high-grade gneisses and pegmatites of the Kerala Khondalite Belt, southern India.
Chemical Geology, 146, 65–85.
Brown, R.L., 2004. Thrust-belt accretion and hinterland underplating of orogenic wedges: an example from the Canadian
Cordillera. In: Thrust Tectonics and Hydrocarbon Systems
(ed. McClay, K.R.), American Association of Petroleum Geologists (AAPG) Memoir 82, 51–64.
Brown, R.L. & Carr, S.D., 1990. Lithospheric thickening and
orogenic collapse within the Canadian Cordillera. Proceedings of the Pacific Rim 90 Congress, Australian Institute of
Mining and Metallogeny, 2, 1–10.
Brown, R.L., McNicoll, V.J., Parrish, R.R. & Scammell, R.J.,
1992. Middle Jurassic plutonism in the Kootenay Terrane,
northern Selkirk Mountains, British Columbia. In: Radiogenic Age and Isotopic Studies: Report 5. Geological Survey
of Canada, Paper 91-2, pp. 135–141.
Brown, S.R., Gibson, H.D., Andrews, G.D.M. et al., 2012.
New constraints on Eocene extension within the Canadian
Cordillera and identification of Phanerozoic protoliths for
footwall gneisses of the Okanagan Valley shear zone. Lithosphere, 4, 354–377.
Caddick, M.J., Bickle, M.J., Harris, N.B.W., Holland, T.J.B.,
Horstwood, M.S.A. & Ahmad, T., 2007. Burial and exhumation history of a Lesser Himalayan schist: recording the formation of an inverted metamorphic sequence in NW India.
Earth and Planetary Science Letters, 264, 375–390.
Carr, S.D., 1991. Three crustal zones in the Thor-Odin-Pinnacles area, southern Omineca belt, British Columbia. Canadian Journal of Earth Sciences, 28, 2003–2023.
Chatterjee, N.D. & Froese, E., 1975. A thermodynamic study
of the pseudobinary join muscovite-paragonite in the system
KAlSi3O8–NaAlSi3O8-Al2O3-SiO2-H2O. American Mineralogist, 60, 985–993.
Cheng, H., Vervoort, J.D., Li, X., Zhang, C., Li, Q. & Zheng,
S., 2011. The growth interval of garnet in the UHP eclogites
from the Dabie orogen, China. American Mineralogist, 96,
1300–1307.
Cherniak, D.J., Watson, E.B., Grove, M. & Harrison, T.M.,
2004. Pb diffusion in monazite: a combined RBS/SIMS
study. Geochimica et Cosmochimica Acta, 68, 829–840.
Cocherie, A., Legendre, O., Peucat, J.J. & Kouamelan, A.N.,
1998. Geochronology of polygenetic monazites constrained
by in situ electron-microprobe Th–U total lead determination–implications for lead behavior in monazite. Geochimica
et Cosmochimica Acta, 62, 2475–2497.
Coggon, R. & Holland, T.J.B., 2002. Mixing properties of
phengitic micas and revised garnet-phengite thermobarometers. Journal of Metamorphic Geology, 20, 683–696.
Colpron, M. & Ryan, J.J., 2010. Bedrock geology of southwest
McQuesten (NTS 115P) and part of northern Carmacks (NTS
115I) map area. In: Yukon Exploration and Geology 2009 (eds
MacFarlane, K.E., Weston, L.H. & Blackburn, L.R.), pp.
159–184. Yukon Geological Survey, Whitehorse, Yukon.
Colpron, M., Price, R.A., Archibald, D.A. & Carmichael,
D.M., 1996. Middle Jurassic exhumation along the western
flank of the Selkirk fan structure: thermobarometric and
thermochronometric constraints from the Illecillewaet synclinorium, southeastern British Columbia. Geological Society of
America Bulletin, 108, 1372–1392.
Colpron, M., Nelson, J.L. & Murphy, D.C., 2006. A tectonostratigraphic framework for the pericratonic terranes of the
northern Canadian Cordillera. In: Paleozoic Evolution and
Metallogeny of Pericratonic Terranes at the Ancient Pacific
Margin of North America, Canadian and Alaskan Cordillera,
Special Paper 45 (eds Colpron, M. & Nelson, J.L.), pp. 1–23.
Geological Association of Canada, St. John’s, Newfoundland.
Colpron, M., Nelson, J.L. & Murphy, D.C., 2007. Northern
Cordilleran terranes and their interactions through time.
GSA Today, 17, 4–10.
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 751
Coney, P.J., 1980. Cordilleran core complexes; an overview.
In: Cordilleran Metamorphic Core Complexes, Memoir 153
(eds Crittenden, M.D., Coney, P.J. & Davis, G.H.), pp. 7–34.
Geological Society of America, Boulder, CO.
Coney, P.J. & Harms, T.A., 1984. Cordilleran metamorphic
core complexes: Cenozoic extensional relics of Mesozoic
compression. Geology, 12, 550–554.
Crowley, J.L. & Ghent, E.D., 1999. An electron microprobe
study of the U-Th-Pb systematics of metamorphosed monazite: the role of Pb diffusion versus overgrowth and recrystallization. Chemical Geology, 157, 285–302.
Crowley, J.L. & Parrish, R.R., 1999. U-Pb isotopic constraints
on diachronous metamorphism in the northern Monashee
complex, southern Canadian Cordillera. Journal of Metamorphic Geology, 17, 483–502.
Crowley, J.L., Ghent, E.D., Carr, S.D., Simony, P.S. & Hamilton, M.A., 2000. Multiple thermotectonic events in a continuous metamorphic sequence, Mica Creek area, southeastern
Canadian Cordillera. Geological Materials Research, 2, 1–45.
Culshaw, N.G., Beaumont, C. & Jamieson, R.A., 2006. The
orogenic superstructure-infrastructure concept: revisited,
quantified, and revived. Geology, 34, 733–736.
De Capitani, C. & Brown, T.H., 1987. The computation of
chemical equilibrium in complex systems containing nonideal solutions. Geochimica et Cosmochimica Acta, 51, 2639–
2652.
De Capitani, C. & Petrakakis, K., 2010. The computation of
equilibrium assemblage diagrams with Theriak/Domino software. American Mineralogist, 95, 1006–1016.
De Sitter, L.U. & Zwart, H.J., 1960. Tectonic development in
supra- and infra-structures of a mountain chain. 21st International Geological Congress, Copenhagen, 18, 249–256.
DeWolf, C.P., Belshaw, N. & O’Nions, R.K., 1993. A metamorphic history from micron-scale 207Pb/206Pb chronometry
of Archean monazite. Earth and Planetary Science Letters,
120, 207–220.
Dusel-Bacon, C., Hansen, V.L. & Scala, J.A., 1995. High-pressure amphibolite facies dynamic metamorphism and the
Mesozoic tectonic evolution of an ancient continental margin, east-central Alaska. Journal of Metamorphic Geology,
13, 9–24.
Dusel-Bacon, C., Lanphere, M.A., Sharp, W.D., Layer, P.W.
& Hansen, V.L., 2002. Mesozoic thermal history and timing
of structural events for the Yukon-Tanana Upland, east-central Alaska: 40Ar/39Ar data from metamorphic and plutonic
rocks. Canadian Journal of Earth Sciences, 39, 1013–1051.
Dusel-Bacon, C., Hopkins, M.J., Mortensen, J.K., Dashevsky,
S.S., Bressler, J.R. & Day, W.C., 2006. Paleozoic tectonic
and metallogenic evolution of the pericratonic rocks of eastcentral Alaska and adjacent Yukon. In: Paleozoic Evolution
and Metallogeny of Pericratonic Terranes at the Ancient Pacific Margin of North America, Canadian and Alaskan Cordillera, Special Paper 45 (eds Colpron, M. & Nelson, J.L.), pp.
1–23. Geological Association of Canada, St. John’s, Newfoundland.
Fitzsimons, I.C.W., Kinny, P.D., Wetherley, S. & Hollingsworth, D.A., 2005. Bulk chemical control on metamorphic
monazite growth in pelitic schists and implications for U-Pb
age data. Journal of Metamorphic Geology, 23, 261–277.
Florence, F.P. & Spear, F.S., 1993. Influences of reaction history and chemical diffusion on P-T calculations for staurolite
schists from the Littleton Formation, northwestern New
Hampshire. American Mineralogist, 78, 345–359.
Foster, G., Gibson, H.D., Parrish, R.R., Horstwood, M., Fraser, J. & Tindle, A., 2002. Textural, chemical and isotopic
insights into the nature and behaviour of metamorphic monazite. Chemical Geology, 191, 183–207.
Foster, G., Parrish, R.R., Horstwood, M.S.A., Chenery, S.,
Pyle, J. & Gibson, H.D., 2004. The generation of prograde
P-T–t points and paths; a textural, compositional, and chronological study of metamorphic monazite. Earth Planetary
Science Letters, 228, 125–142.
© 2013 John Wiley & Sons Ltd
Fuhrman, M.L. & Lindsley, D.H., 1988. Ternary-feldspar modeling and thermometry. American Mineralogist, 73, 201–216.
Gabrielse, H., Murphy, D.C. & Mortensen, J.K., 2006. Cretaceous and Cenozoic dextral orogen-parallel displacements,
magmatism and paleogeography, north-central Canadian
Cordillera. In: Paleogeography of the North American Cordillera: Evidence for and Against Large-Scale Displacements, Special Paper 46 (eds Haggart, J.W., Monger, J.W.H. & Enkin,
R.J.), pp. 255–276. Geological Association of Canada, St.
John’s, Newfoundland.
Gaidies, F., Krenn, E., de Capitani, C. & Abart, R., 2008.
Coupling forward modelling of garnet growth with monazite
geochronology: an application to the Rappold complex
(Austroalpine crystalline basement). Journal of Metamorphic
Geology, 26, 775–793.
Gardes, E., Jaoul, O., Montel, J.-M., Seydoux-Guillaume,
A.-M. & Wirth, R., 2006. Pb diffusion in monazite: an
experimental study of Pb2+ +Th4+ , 2Nd3+ interdiffusion.
Geochimica et Cosmochimica Acta, 70, 2325–2336.
Gibson, H.D., Brown, R.L. & Parrish, R.R., 1999. Deformation-induced inverted metamorphic field gradients: an example from the southeastern Canadian Cordillera. Journal of
Structural Geology, 21, 751–767.
Gibson, H.D., Carr, S.D., Brown, R.L. & Hamilton, M.A.,
2004. Correlations between chemical and age domains in
monazite, and metamorphic reactions involving major politic
phases: an integration of ID-TIMS and SHRIMP geochronology with Y-Th–U X-ray mapping. Chemical Geology,
211, 237–260.
Gibson, H.D., Brown, R.L. & Carr, S.D., 2005. U-Th- Pb geochronologic constraints on the structural evolution of the
Selkirk fan, northern Selkirk Mountains, southern Canadian
Cordillera. Journal of Structural Geology, 27, 1899–1924.
Gibson, H.D., Brown, R.L. & Carr, S.D., 2008. Tectonic evolution of the Selkirk fan, southeastern Canadian Cordillera:
a composite Middle Jurassic–Cretaceous orogenic structure.
Tectonics, 27, TC6007, doi: 10.1029/2007TC002160.
Goncalves, P., Williams, M.L. & Jercinovic, M.J., 2005. Electron-microprobe age mapping of monazite. American Mineralogist, 90, 578–585.
Gordey, S.P. & Ryan, J.J., 2005. Geology, Stewart River area
(115N, 115-O and part of 115J), Yukon Territory. Geological Survey of Canada, Open File 4970 (1 sheet, 1:250 000
scale).
Hansen, V.L. & Dusel-Bacon, C., 1998. Structural and Kinematic evolution of the Yukon-Tanana upland tectonites,
east-central Alaska: a record of late Paleozoic to Mesozoic
crustal assembly. Geological Society of America Bulletin, 110,
211–230.
Hansen, V.L., Heizler, M.T. & Harrison, T.M., 1991. Mesozoic
thermal evolution of the Yukon-Tanana composite terrane;
new evidence from 40Ar/39Ar data. Tectonics, 10, 51–76.
Harlov, D.E. & Hetherington, C.J., 2010. Partial high-grade
alteration of monazite using alkali-bearing fluids: experiment
and nature. American Mineralogist, 95, 1105–1108.
Harlov, D.E., Wirth, R. & Hetherington, C.J., 2011.
Fluid-mediated partial alteration in monazite: the role of
coupled dissolution–reprecipitation in element redistribution
and mass transfer. Contributions to Mineralogy and Petrology, 162, 329–348.
Hayward, N., Miles, W. & Oneschuk, D. 2012. Geophysical
Series, detailed geophysical compilation project, Yukon Plateau, Yukon, NTS 115-I, J, K, N, O, P and 116A and B.
Geological Survey of Canada, Open File 7279 (2 sheets,
1:350 000 Scale).
Hetherington, C.J., Harlov, D.E. & Budzy
n, B., 2010. Experimental metasomatism of monazite and xenotime: mineral
stability, REE mobility and fluid composition. Mineralogy
and Petrology, 99, 165–184.
Holland, T.J.B. & Powell, R., 1998. An internally consistent
thermodynamic data set for phases of petrological interest.
Journal of Metamorphic Geology, 16, 309–343.
752 R. D. STAPL ES ET AL .
Holland, T. & Powell, R., 2003. Activity-composition relations
for phases in petrological calculations; an asymmetric multicomponent formulation. Contributions to Mineralogy and
Petrology, 145, 492–501.
Horvath, P., Balen, D., Finger, F., Tomljenovic, B. & Krenn,
E., 2010. Contrasting P-T–t paths from the basement of the
Tisia Unit (Slavonian Mts., NE Croatia): application of
quantitative phase diagrams and monazite age dating.
Lithos, 117, 269–282.
Hunt, P.A. & Roddick, J.C., 1992. A compilation of K–Ar
and 40Ar–39Ar ages: report 22. In: Radiogenic Age and Isotopic Studies: Report 6. Geological Survey of Canada, Paper
92-2, pp. 179–226.
Ireland, T.R. & Gibson, G.M., 1998. SHRIMP monazite and
zircon geochronology of high-grade metamorphism in New
Zealand. Journal of Metamorphic Geology, 16, 149–167.
Janots, E., Engi, M., Berger, A., Allaz, J., Schwarz, J.O. &
Spandler, C., 2008. Prograde metamorphic sequence of REE
minerals in pelitic rocks of the Central Alps, implications for
allanite–monazite–xenotime phase relations from 250 to
610 °C. Journal of Metamorphic Geology, 26, 509–526.
Johnston, S.T., Mortensen, J.K. & Erdmer, P., 1996. Igneous
and metaigneous age constraints for the Aishihik metamorphic suite, southwest Yukon. Canadian Journal of Earth Sciences, 33, 1543–1555.
Knight, E., Schneider, D.A. & Ryan, J.J., 2013. Thermochronology of the Yukon-Tanana terrane, west-central Yukon:
evidence for Jurassic extension and exhumation in the northern Canadian Cordillera. Journal of Geology, 121, 371–400
Kohn, M.J. & Spear, F., 2000. Retrograde net transfer reaction insurance for pressure-temperature estimates. Geology,
28, 1127–1130.
Kretz, R., 1983. Symbols for rock-forming minerals. American
Mineralogist, 68, 277–279.
Lowey, G.W. & Hills, L.V., 1988. Lithofacies, petrography,
and environments of deposition, Tantalus Formation (Lower
Cretaceous), Indian River area, west-central Yukon. Bulletin
of Canadian Petroleum Geology, 36, 296–310.
Ludwig, K.R., 2008. Manual for Isoplot 3.7: A Geochronological Toolkit for Microsoft Excel. Special Publication No. 4.
rev. August 26, 2008, Berkeley Geochronology Center, Berkeley, CA, 77pp.
Mackenzie, D. & Craw, D., 2012. Contrasting structural settings of mafic and ultramafic rocks in the Yukon-Tanana
terrane. In: Yukon Exploration and Geology 2011 (eds MacFarlane, K.E. & Sack, P.J.), pp. 115–127. Yukon Geological
Survey, Whitehorse, Yukon.
Mattison, C.G., Wooden, J.L., Liou, J.G., Bird, D.K. & Wu,
C.L., 2006. Age and duration of eclogite-facies metamorphism, north Qaidam HP/UHP terrane, western China.
American Journal of Science, 306, 683–711.
Monger, J.W.H. & Price, R.A., 2002. The Canadian Cordillera: geology and tectonic evolution. Canadian Society of
Exploration Geophysicists Recorder, 27, 17–36.
Monger, J.W.H., Price, R.A. & Tempelman-Kluit, D.J., 1982.
Tectonic accretion and the origin of the two major metamorphic and plutonic welts in the Canadian Cordillera. Geology,
10, 70–75.
Mortensen, J.K., 1990. Geology and U-Pb geochronology of
the Klondike District, west-central Yukon Territory. Canadian Journal of Earth Sciences, 27, 903–914.
Murphy, D.C., 2004. Devonian-Mississippian metavolcanic
stratigraphy, massive sulphide potential and structural
re-interpretation of Yukon-Tanana Terrane south of the Finlayson Lake massive district, southeastern Yukon (105G/1,
105H/3,4,5). In: Yukon Exploration and Geology 2003 (eds
Emond, D.S. & Lewis, L.L.), pp. 157–175. Yukon Geological Survey, Whitehorse, Yukon.
Murphy, D.C., van der Heyden, P., Parrish, R.R. et al., 1995.
New geochronological constraints on Jurassic deformation
of the western edge of North America, southeastern Canadian Cordillera. In: Jurassic Magmatism and Tectonics of the
North American Cordillera, Special Paper 299 (eds Miller,
D.M. & Busby, C.), pp. 159–171. Geological Society of
America, Boulder, CO.
Nelson, J.L., Colpron, M., Piercey, S.J., Dusel-Bacon, C., Murphy, D.C. & Roots, C.F., 2006. Paleozoic tectonic and metallogenic evolution of the pericratonic terranes in Yukon,
northern British Columbia and eastern Alaska. In: Paleozoic
Evolution and Metallogeny of Pericratonic Terranes at the
Ancient Pacific Margin of North America, Canadian and Alaskan Cordillera, Special Paper 45 (eds Colpron, M. & Nelson,
J.L.), pp. 323–360. Geological Association of Canada, St.
John’s, Newfoundland.
Parrish, R.R., 1995. Thermal evolution of the southeastern
Canadian Cordillera. Canadian Journal of Earth Sciences, 32,
1618–1642.
Parrish, R.R., Carr, S.D. & Parkinson, D.L., 1988. Eocene
extensional tectonics and geochronology of the southern
Omineca belt, British Columbia and Washington. Tectonics,
7, 181–212.
Pattison, D.R.M. & Tinkham, D.T., 2009. Interplay between
equilibrium and kinetics in prograde metamorphism of pelites: an example from the Nelson aureole, British Columbia.
Journal of Metamorphic Geology, 27, 249–279.
Pattison, D.R.M., de Capitani, C. & Gaidies, F., 2011. Petrological consequences of variations in metamorphic reaction
history. Journal of Metamorphic Geology, 29, 953–977.
Pavlis, T.L., Sisson, V.B., Foster, H.L., Nockleberg, W.J. &
Plafker, G., 1993. Mid-Cretaceous extensional tectonics of
the Yukon–Tanana terrane, Trans-Alaskan Crustal Transect
(TACT), east-central Alaska. Tectonics, 12, 103–122.
Pyle, J.M. & Spear, F.S., 1999. Yttrium zoning in garnet: coupling of major and accessory phases during metamorphic
reactions. Geological Materials Research, 1, 1–49.
Rayner, N. & Stern, R.A., 2002. Improved Sample Preparation
Method for SHRIMP Analysis of Delicate Mineral Grains
Exposed in Thin Sections. Geological Survey of Canada,
Current Research 2002-F10, 1–3.
Ring, U., Brandon, M.T., Willet, S.D. & Lister, G.S., 1999.
Exhumation processes. In: Exhumation Processes: Normal
Faulting, Ductile Flow, and Erosion (eds Ring, U., Brandon,
M.T., Lister, G.S. & Willet, S.D.), pp. 1–27. Geological
Society of London, Special Publications.
Ruks, T.W., Piercey, S.J., Ryan, J.J., Villeneuve, M.E. & Creaser, R.A., 2006. Mid- to Late Paleozoic K-feldspar augen
granitoids of the Yukon-Tanana terrane, Yukon: implications for crustal growth and tectonic evolution of the northern Cordillera. Geological Society of America Bulletin, 118,
1212–1231.
Ryan, J.J., Gordey, S.P., Glombick, P., Piercey, S.J. & Villeneuve, M.E., 2003. Update on Bedrock Geological Mapping
of the Yukon-Tanana Terrane, Southern Stewart River Map
Area, Yukon Territory. Geological Survey of Canada,
Current Research 2003-A9. 7p.
Ryan, J.J., Colpron, M. & Hayward, N., 2010. Geology,
southwestern McQuesten and parts of northern Carmacks,
Yukon; Geological Survey of Canada, Canadian Geoscience
Map 7, (preliminary version), scale 1:125 000.
Sch€
arer, U., 1984. The effect of initial 230Th disequilibrium on
young U-Pb ages: the Makalu case, Himalaya. Earth and
Planetary Science Letters, 67, 191–204.
Seydoux-Guillaume, A.M., Paquette, J.L., Wiedenbeck, M.,
Montel, J.-M. & Heinrich, W., 2002. Experimental resetting
of the U-Th–Pb systems in monazite. Chemical Geology, 191,
165–181.
Skora, S., Lapen, T.J., Baumgartner, L.P., Johnson, C.M., Hellebrand, E. & Mahlen, M.J., 2009. The duration of prograde
garnet crystallization in the UHP eclogites at Lago di
Cignana, Italy. Earth and Planetary Science Letters, 287, 402–
411.
Spear, F.S. & Parrish, R.R., 1996. Petrology and cooling rates
of the Valhalla complex, British Columbia, Canada. Journal
of Petrology, 37, 733–765.
© 2013 John Wiley & Sons Ltd
METAMO RPHISM WITHIN YUKON-TANANA TERRANE 753
Spear, F.S., Kohn, M.J., Florence, F.P. & Menard, T., 1990.
A model for garnet and plagioclase growth in pelitic schists:
implications for thermobarometry and P-T path determinations. Journal of Metamorphic Geology, 8, 683–696.
Stacey, J.S. & Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a two-stage model. Earth and
Planetary Science Letters, 26, 207–221.
Stern, R.A., 1997. The GSC sensitive high resolution ion microphobe (SHRIMP): analytical techniques of zircon
U-Th-Pb age determinations and performance evaluation.
Radiogenic Age and Isotopic Studies: Report 10. Geological
Survey of Canada, Current Research 1997-F, 1–31.
Stern, R.A. & Berman, R.G., 2000. Monazite U-Pb and Th-Pb
geochronology by ion microprobe, with an application to in
situ dating of an Archean metasedimentary rock. Chemical
Geology, 172, 113–130.
Stern, R.A. & Sanborn, N., 1998. Monazite U–Pb and Th–Pb
geochronology by high-resolution secondary ion mass spectrometry. Radiogenic Age and Isotopic Studies: Report 11.
Geological Survey of Canada, Current Research 1998-F. 1–18.
Stevens, R.A., Mortensen, J.K. & Hunt, P.A., 1993. U–Pb and
40Ar/39Ar geochronology of plutonic rocks from the Teslin
suture zone, Yukon Territory. In: Radiogenic and Isotope
Studies, Report 7. Geological Survey of Canada, Paper 93-2,
pp. 83–90.
Struik, L.C., 1993. Intersecting intracontinental Tertiary transform fault systems in the North American Cordillera. Canadian Journal of Earth Sciences, 30, 1262–1274.
St€
uwe, K., 1997. Effective bulk composition changes due to
cooling: a model predicting complexities in retrograde reaction
textures. Contributions to Mineralogy and Petrology, 129,
43–52.
Tinkham, D.K., Zuluaga, C.A. & Stowell, H.H., 2001. Metapelite phase equilibria modeling in MnNCKFMASH: the
effect of variable Al2O3 and MgO/(MgO+FeO) on mineral
stability. Geological Materials Research, 3, 1–42.
Tobisch, O.T. & Paterson, S.C., 1988. Analysis and interpretation of composite foliations in an area of progressive deformation. Journal of Structural Geology, 10, 745–754.
Todd, C., 1998. Limits on the precision of geobarometry at
low grossular and anorthite content. American Mineralogist,
83, 1161–1167.
Tracy, R.J., Robinson, P. & Thompson, A.B., 1976. Garnet
composition and zoning in the determination of temperature
and pressure of metamorphism, central Massachusetts.
American Mineralogist, 61, 762–775.
Vance, D. & Mahar, E., 1998. Pressure–temperature paths
from P-T pseudosections and zoned garnets; potential, limitations and examples from the Zanskar Himalaya, NW
India. Contributions to Mineralogy and Petrology, 132, 225–
245.
Waters, D.J. & Lovegrove, D.P., 2002. Assessing the extent
of disequilibrium and overstepping of prograde metamor-
© 2013 John Wiley & Sons Ltd
phic reactions in metapelites from the Bushveld Complex
aureole, South Africa. Journal of Metamorphic Geology, 20,
135–149.
White, R.W., Powell, R., Holland, T.J.B. & Worley, B.A.,
2000. The effect of TiO2 and Fe2O3 on metapelitic assemblages at greenschist and amphibolite facies conditions:
mineral equilibria calculations in the system K2O–FeO–MgO–
Al2O3–SiO2–H2O–TiO2–Fe2O3. Journal of Metamorphic
Geology, 18, 497–511.
White, R.W., Powell, R. & Holland, T.J.B., 2007. Progress
relating to calculations of partial melting equilibria for
metapelites. Journal of Metamorphic Geology, 25, 511–527.
Williams, P.F., 1983. Large scale transposition by folding in
Northern Norway. Geologische Rundschau, 72, 589–604.
Williams, P.F., 1985. Multiply deformed terrains-problems of
correlation. Journal of Structural Geology, 7, 269–280.
Williams, P.F. & Compagnoni, R., 1983. Deformation and
metamorphism in the Bard area of the Sesia Lanzo zone,
Western Alps, during subduction and uplift. Journal of
Metamorphic Geology, 1, 117–140.
Williams, M.L. & Jercinovic, M.J., 2002. Microprobe monazite
geochronology: putting absolute time into microstructural
analysis. Journal of Structural Geology, 24, 1013–1028.
Williams, M.L., Jercinovic, M.J. & Terry, M.P., 1999. Age
mapping and dating of monazite on the electron microprobe:
Deconvoluting multistage tectonic histories. Geology, 27,
1023–1026.
Williams, M.L., Jercinovic, M.J., Harlov, D.E., Budzy
n, B. &
Hetherington, C.J., 2011. Resetting monazite ages during
fluid-related alteration. Chemical Geology, 283, 218–225.
Wolf, D.E., Andronicos, C.L., Vervoort, J.D., Mansfield,
M.R. & Chardon, D., 2010. Application of Lu-Hf garnet
dating to unravel the relationships between deformation,
metamorphism and plutonism: an example from the Prince
Rupert area, British Columbia. Tectonophisics, 485, 62–77.
Zhu, X.K. & O’Nions, R.K., 1999. Zonation of monazite in
metamorphic rocks and its implications for high temperature
thermochronology: a case study from the Lewisian terrain.
Earth and Planetary Science Letters, 171, 209–220.
SUPPORTING INFORMATION
Additional Supporting Information may be found in
the online version of this article at the publisher’s
web site:
Table S1. SHRIMP U-Th-Pb analytical data for
monazite corrected using the 204Pb-method.
Received 17 January 2013; revision accepted 24 May 2013.
Download